Vous êtes sur la page 1sur 23

Analytica Chimica Acta 890 (2015) 60e82

Contents lists available at ScienceDirect

Analytica Chimica Acta


journal homepage: www.elsevier.com/locate/aca

Review

The molybdenum blue reaction for the determination of


orthophosphate revisited: Opening the black box
Edward A. Nagul a, b, Ian D. McKelvie a, c, Paul Worsfold c, Spas D. Kolev a, b, *
a
School of Chemistry, The University of Melbourne, Victoria 3010, Australia
b
Centre for Aquatic Pollution Identification and Management (CAPIM), The University of Melbourne, Victoria 3010, Australia
c
School of Geography, Earth and Environmental Sciences, Plymouth University, Plymouth PL48AA, UK

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 Molybdenum blue chemistry for


orthophosphate determination is
discussed.
 The choice of reductant determines
the blue product(s) obtained.
 Mechanisms are described for
various additive and subtractive
interferents.
 The choice of strong mineral acid for
the reaction should be considered.
 Detailed recommendations are made
for method optimisation.

a r t i c l e i n f o a b s t r a c t

Article history: The molybdenum blue reaction, used predominantly for the determination of orthophosphate in envi-
Received 6 May 2015 ronmental waters, has been perpetually modified and re-optimised over the years, but this important
Received in revised form reaction in analytical chemistry is usually treated as something of a 'black box' in the analytical literature.
19 July 2015
A large number of papers describe a wide variety of reaction conditions and apparently different
Accepted 23 July 2015
Available online 10 August 2015
products (as determined by UVevisible spectroscopy) but a discussion of the chemistry underlying this
behaviour is often addressed superficially or not at all. This review aims to rationalise the findings of the
many 'optimised' molybdenum blue methods in the literature, mainly for environmental waters, in terms
Keywords:
Molybdenum blue reaction
of the underlying polyoxometallate chemistry and offers suggestions for the further enhancement of this
Orthophosphate time-honoured analytical reaction.
Dissolved reactive phosphorus © 2015 Elsevier B.V. All rights reserved.
Phosphomolybdate

List of abbreviations: (br), Broad absorption band; (sh), Absorption shoulder; 12-MPA, 12-Molybdophosphoric acid (H3PMo12O40); 11-MPA, 11-Molybdophosphoric acid
(H3PMo11O37); 12-MSA, 12-Molybdosilicic acid (H4SiMo12O40); AA, Ascorbic acid; ANS, 1-Amino-2-naphthol-4-sulfonic acid; AsMB, Arsenomolybdenum blue; DA, Discrete
analyser; DAPH, 2,4-Diaminophenol dihydrochloride; DOP, Dissolved organic phosphorus; MB, Molybdenum blue; DRP, Dissolved reactive phosphorus; ESI-MS, Electrospray
ionisation mass spectrometry; FIA, Flow injection analysis; HQ, Hydroquinone; HS, Hydrazine sulfate; IVCT, Intervalence charge transfer; LMCT, Ligand-metal charge transfer;
Metol, 4-(Methylamino)phenol sulfate; MRP, Molybdate reactive phosphorus; PMB, Phosphomolybdenum blue; rFIA, Reverse flow injection analysis; SFA, Segmented
continuous flow analysis; SIA, Sequential injection analysis; SiMB, Silicomolybdenum blue; TDP, Total dissolved phosphorus.
* Corresponding author. School of Chemistry, The University of Melbourne, Victoria 3010, Australia.
E-mail address: s.kolev@unimelb.edu.au (S.D. Kolev).

http://dx.doi.org/10.1016/j.aca.2015.07.030
0003-2670/© 2015 Elsevier B.V. All rights reserved.
E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82 61

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
2. Chemistry of the molybdenum blue reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
2.1. Reaction overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
2.2. Mo(VI) speciation and 12-MPA formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
2.3. Redox chemistry of PMB . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
2.3.1. Reduction of 12-MPA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
2.3.2. Spectral features of PMB . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
2.3.3. Nature of the reduced products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
2.3.4. Isomerism of 12-MPA and its reduced forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
2.3.5. Organic reductants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
2.3.6. Metallic reductants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
3. MB method optimisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.1. Reagent concentrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.2. The reagent blank: isopoly molybdenum blue species . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
3.3. Product stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
3.4. Method linearity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
3.5. Choice of acid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4. Interferences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.1. Additive interferences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.1.1. Arsenate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.1.2. Silicate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.1.3. Organic and inorganic P hydrolysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.2. Subtractive interferences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.2.1. Organic acids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.2.2. Fluoride . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.2.3. Chloride (salt error) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.3. Multifunctional interferents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.3.1. Sulfide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.3.2. Iron . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.3.3. Surfactants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
5. MB chemistry in flow methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
6. Conclusions and recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
6.1. Recommended reductants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
6.2. Recommended acids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
6.3. Recommended optimisation procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

1. Introduction (MB) reaction is the most common means of determination [1]. It


can also be used for the spectrophotometric determination of sili-
Orthophosphate concentration is a key water quality parameter cate, arsenate and germanate. Strictly, this reaction determines the
and spectrophotometric detection using the molybdenum blue 'molybdate reactive phosphorus' (MRP) fraction which includes
other labile phosphorus species in addition to orthophosphate [2]
as discussed in Section 4.1.
The reaction involves the formation of a polyoxometallate spe-
cies, a heteropoly acid, from orthophosphate and molybdate under
acidic conditions, which is then reduced to form an intensely col-
oured phosphomolybdenum blue (PMB) species. This reaction was
mentioned by Scheele in 1783, but its discovery is widely attributed
to Berzelius (1826) [3]. It was not until 1934, however, that Keggin
proposed the structures of a range of 12-heteropoly acids [4]
(Fig. 1). ‘Molybdenum blue’ refers not to a single species, but
rather to a family of reduced molybdate compounds, which may or
may not contain a heteroatom, e.g. phosphorus. Distinction be-
tween heteropoly (containing a hetero-atom) and isopoly (con-
taining no hetero-atom) molybdenum blue species is made in this
review where necessary.
A fundamental understanding of the inorganic chemistry of the
MB reaction is important for optimising its analytical application
for the determination of orthophosphate. In particular, the con-
Fig. 1. Structure of the Keggin ion [PW12O40]3, analogous to that of [PMo12O40]3. The centrations of the reagents can be optimised to maximise the de-
black, grey and white spheres represent P, W and O respectively. Reproduced from gree of product formation and product stability (for batch methods)
Ref. [5] with permission from The Royal Society of Chemistry.
62 E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82

and achieve good precision and accuracy. Table 1


This review systematically summarises the fundamental Predicted pH values for maximum concentrations of Mo(VI) species* in 10 mmol L1
Mo(VI) solution. Shorthand (dehydrated) forms for molybdic acid and the two cat-
chemistry of the MB reaction and discusses the optimal conditions ions are also shown.
for the selective determination of MRP using batch methods under
equilibrium conditions. The additional requirements for non- Species Name Z pH

equilibrium, flow-based methods are also considered. MoO4 2


Molybdate 0.00 7.00
HMoO4  H-Molybdate 1.00 4.58
Mo2 O7 2 Dimolybdate 1.00 4.58
2. Chemistry of the molybdenum blue reaction Mo7 O24 6 Heptamolybdate 1.14 4.30
HMo7 O24 5 H-Heptamolybdate 1.29 3.64
2.1. Reaction overview Mo3 O10 2 Trimolybdate 1.33 3.52
H2 Mo7 O24 4 H2-Heptamolybdate 1.43 2.85
Mo4 O13 2 Tetramolybdate 1.50 2.25
The MB reaction occurs in two stages; the first stage involves the HMo2 O7  H-Dimolybdate 1.50 2.25
formation of a Keggin ion around the analyte anion and the second Mo8 O26 4 Octamolybdate 1.50 2.25
stage entails the reduction of this heteropoly acid to form a deeply HMo3 O10  H-Trimolybdate 1.67 1.48
blue-coloured product. These stages can be described in the Mo6 O19 2 Hexamolybdate 1.67 1.48
HMo4 O13  H-Tetramolybdate 1.75 1.33
simplified forms shown in Eqs. (1) and (2) [2]. HMo5 O16  H-Pentamolybdate 1.80 1.28
Mo11 O34 2 Undecamolybdate 1.82 1.27
þ
PO3 2
4 þ 12MoO4 þ 27H /H3 PO4 ðMoO3 Þ12 þ 12H2 O (1) HMo6 O19  H-Hexamolybdate 1.83 1.27
Mo12 O37 2 Dodecamolybdate 1.83 1.25
 3 MoO3 or MoO3(OH2)3 Molybdic acid 2.00 1.20
H3 PMoðVIÞ12 O40 þ Reductant/ H4 PMoðVIÞ8 MoðVÞ4 O40 HMo2 O6 þ or Mo2O(OH)9(OH2)þ Molybdate dimer cation 2.50 <1.20
HMoO3 þ or MoO2(OH)ðOH2 Þþ
3 Molybdate monocation 3.00 <1.20
(2)
*
Species and Z values are drawn from Refs. [6e8,10e20].
All MB methods require a strong acid, a source of Mo(VI) and a
reductant, normally in aqueous solution. The concentrations of acid
and molybdate are vital, not only for the formation of the heter- (½Hþ þ þ þ
total ) involving both free H (½Hfree ) and H bound to proton-
opoly acid but also for controlling its reduction. It is well-known ated and/or condensed Mo(VI) must be considered. In addition to
that orthophosphate (PO4 3 ), like other tetrahedral anions of the the definition above, Z is also expressed in terms of the concen-
form XO4, form Keggin ions of composition [XnþMo12O40](8n) trations of the relevant species in solution (Eq. (4)), and has been
where X is the heteroatom. In the case of orthophosphate, this determined for a wide range of solution compositions (Fig. 2).
species is known as 12-molybdophosphoric acid (12-MPA). It is also
well-established that procedures for orthophosphate determina- ½Hþ þ
total  ½Hfree
Z¼ (4)
tion are best carried out at pH 0e1; the reasons for this are seldom ½MoðVIÞ
explained in terms of the molybdate chemistry but are empirically
The molybdate anion can be protonated twice to form neutral
selected to give the best colour intensity. Many reductants and
molybdic acid (Eqs. (5)e(6)) [21].
reaction conditions have been used in the MB reaction, typically
giving rise to a mixture of reduced products as evidenced by the þ 
wide variety of absorption maxima and molar absorptivities MoO2
4 þ H #½HMoO4  pKa ¼ 3:66 (5)
reported.
½HMoO4  þ Hþ þ 2H2 O#MoO3 ðOH2 Þ3 pKa ¼ 3:81 (6)
2.2. Mo(VI) speciation and 12-MPA formation When the solution is sufficiently acidic, the second protonation
also involves a change in the coordination chemistry of Mo(VI) from
An understanding of acidic molybdate speciation and the the 4-coordinate HMoO4 2 to the 6-coordinate hydrated molybdic
mechanism of heteropoly acid formation in the MB reaction is acid MoO3(OH2)3, which significantly decreases both of its pKa
necessary to optimise the practical application of this reaction. Key values [13,22]. The process of hydrating molybdic acid in the sec-
factors are the acid and molybdate concentrations, sources of ond protonation step requires that the acid concentration be at
interference and the range of chemical conditions under which this least ten-fold in excess of that of Mo(VI), otherwise precipitation of
reaction is useable. The MB reaction for orthophosphate is usually
performed between pH 0e1, as this range appears to be necessary
to form suitable amounts of a stable reduced product without 2.0
excessive direct reduction of Mo(VI). However, it is not pH alone 1.8
that determines Mo(VI) speciation; rather, it is a combination of 1.6
acid concentration and molybdate concentration. These parameters 1.4 0.1000 M

determine the ‘degree of protonation’, Z, which is defined as the 1.2 0.0500 M

average number of protons bound to molybdate in solution [6e9]. Z 0.0250 M


1.0
Z

0.0100 M
reflects the complex solution equilibria of Mo(VI); at a given Mo(VI) 0.8
0.0050 M
concentration, the Z value of a solution can be used to predict the 0.6
0.0025 M
predominant molybdate species present as a function of pH 0.4
0.0010 M
(Table 1). Z is defined as the ratio q/p in Eq. (3). 0.2
0.0005 M
0.0
  h ið2pqÞ
þ qHþ # ðMoO4 Þp Hq
0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0
p MoO2
4 (3)
pH
Using this ratio is not as simple as using the often cited [Hþ]/ Fig. 2. Plots of the degree of protonation, Z, vs. pH at different Mo(VI) concentrations
[Mo(VI)] ratio; but the latter is generally misleading as discussed in in 1 mol L1 NaCl to reduce the influence of ionic strength variation. Reproduced from
Section 3.1. Instead, the total proton concentration in solution Ref. [6] with permission from Elsevier.
E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82 63

poorly soluble MoO3 may occur [23]. However, molybdic acid will although these authors also allowed for the possibility of a pre-
begin to protonate and dimerise when Z > 2 (Eqs. (7)e(9)) equilibrium between the dimers and a larger Mo(VI) ion which
[8,11,16,17,24,25], which is undesirable for 12-MPA formation. constituted the actual reactive species. More recent UVevisible
spectrophotometry and Raman spectroscopy studies [7,8,10,32]
 þ
MoO3 ðOH2 Þ3 þ Hþ # MoO2 ðOHÞðOH2 Þ3 pKa z0:9 (7) have shown that this latter consideration was in fact much more
accurate than the theory of cationic dimers as the reactive species,
as it accounts for the decomposition of 12-MPA as well as the in-
 þ  2þ
2 MoO2 ðOHÞðOH2 Þ3 # Mo2 OðOHÞ8 ðOH2 Þ2 þ H2 O hibition of its formation at high acidities where dimeric Mo(VI)
cations actually predominate. In fact, dodecamolybdate
Kd ¼ 100
(Mo12 O37 2 ) appears to be the main precursor Mo(VI) species in
(8) equilibrium with 12-MPA [7,8,10], and a detailed mechanism of 12-
MPA formation is given later in this section. The prevalence of the
 2þ  þ dodecamolybdate anion at very low pH values is minimal since the
Mo2 OðOHÞ8 ðOH2 Þ2 H Mo2 OðOHÞ9 ðOH2 Þ2 þ Hþ
(9) condensation of cationic Mo(VI) species into larger ions such as
pKa z0:1e0:2 Mo12 O37 2 requires the release of Hþ into an already highly acidic
In order for acidified Mo(VI) to polymerise, which is a prereq- solution, as per Eq. (10). Since high Z solutions are unfavourable for
uisite for the formation of MB, Mo(VI) concentrations of at least the existence of the 12-MPA precursor species, 12-MPA dissociates
103 e 102 mol L1 are required [24]. The prevalent species in under these conditions.
solution at various Z values, assuming 10 mmol L1 Mo(VI), are
 þ
shown in Table 1, whilst the relationship between pH, [Mo(VI)] and 6 Mo2 OðOHÞ9 ðOH2 Þ % Mo12 O37 2 þ 29H2 O þ 8Hþ
Z is shown in Fig. 2, and the Mo(VI) solution equilibria are shown in
(10)
Fig. 3. The low Mo(VI) concentrations and ionic strengths used in
the MB reaction actually simplify matters, as larger polymolybdates Another important consideration is the [Mo(VI)]/[P] ratio used
tend to form when either of these parameters is considerably in some studies of phosphomolybdate speciation. When [Mo(VI)]/
higher [6,14,26]. It should also be noted that the choice of Na2MoO4 [P] < 12, the speciation of phosphomolybdates as a function of Z
or (NH4)6Mo7O24.7H2O for MB methods is irrelevant due to the changes markedly, yet it has previously been presumed that
equilibria shown in Fig. 3, provided that sufficient time is allowed speciation under these conditions is true of all P e Mo(VI) systems
for the equilibration of Mo(VI) species before the solution is used. [33]. In MB methods, Mo(VI) is present in large excess over P, and
Whilst it has been reported that several hours are needed for this these Mo(VI)-deficient conditions do not apply.
process in pure solutions of either Na2MoO4 or (NH4)6Mo7O24.4H2O It has been shown via Raman spectroscopy that each anionic
[27], equilibration appears to occur much more quickly if the so- molybdate species in the presence of orthophosphate favours the
lution is pre-mixed with acid, requiring only about 10e30 min formation of one particular molybdophosphoric acid [8] (Fig. 3)
[26,28]. Furthermore, Na2MoO4 may be the favoured Mo(VI) salt in where [Mo(VI)]/[P]  12, with mixtures of Mo(VI) species leading to
laboratories where P and NH3 determinations are carried out mixtures of heteropoly acids. For example, both 12-MPA and 11-
simultaneously, in order to avoid cross-contamination. MPA are formed effectively where molybdic acid, dodeca-
It is important to consider how 12-MPA is formed from its molybdate and hexamolybdate are prevalent [32] (Fig. 4). It is
precursors, as a number of studies addressing this question are therefore important to use reaction mixtures with Z z 2, or
based on several early conclusions made before more detailed data pH  0.9 assuming a Mo(VI) concentration of 10 mmol L1, to
about Mo(VI) equilibria were available in the literature. For ensure orthophosphate is present only as 12-MPA and not as a
example, explanations of both pH-related behaviour [29] and the mixture of 12-MPA and its hydrolysis product, 11-MPA. 11-MPA is
observed rate laws for 12-MPA formation [23,27,30,31] were very difficult to reduce [33], and the molar absorptivity of its 4e-
initially based on conclusions about Mo(VI) and Hþ stoichiometry reduced form is probably quite low owing to its low molecular
obtained under highly acidic conditions [27] which favoured symmetry [34].
monomeric and dimeric Mo(VI) cations (Table 1), and were not Two mechanisms for 12-MPA formation have been proposed
optimal for 12-MPA formation. It was suggested in these studies based on the known equilibria of Mo(VI) and time-resolved Raman
that the assembly of 12-MPA occurred via the reaction of cationic spectroscopy measurements of individual Mo(VI) species [10]. The
Mo(VI) dimers with H3PO4 and the release of Hþ [23,27,30,31], first mechanism (Eq. (11)), known as the 'displacement'

MoO42- HMoO4- Mo7O246- HMo7O245- Mo8O264- Mo6O192- Mo12O372- MoO3(OH2)3 Mo2O(OH)9(OH2)+

pH 7.00 4.58 4.37 4.10 3.72 3.16 2.25 1.65 1.40 1.28 1.22 1.20 1.05 1.00 0.95 0.91 0.87

Z 0.00 1.00 1.10 1.20 1.30 1.40 1.50 1.60 1.70 1.80 1.90 2.00 2.10 2.20 2.30 2.40 2.50

PMo12O403-
P2Mo5O236-

PMo11O373-

Fig. 3. Mo(VI) speciation in aqueous solution containing 10 mmol L1 Mo(VI) as a function of Z (bottom), or as a function of pH (top), and the phosphomolybdate species which may
form under these conditions (deprotonated forms shown for clarity). The optimal Z value for the formation of each phosphomolybdate and Mo(VI) species is indicated by a vertical
line. The width of the box for each phosphomolybdate denotes the Z range in which it has been observed; a dashed line indicates that the prevalence of the complex at higher Z
values has not been clearly characterised. Note that pH values for Z > 2 have been extrapolated from literature data (Fig. 2) [6], and represent approximations only. Phospho-
molybdate speciation data adapted from Ref. [8] with permission from Elsevier.
64 E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82

mechanism, occurs when orthophosphate is added to Mo(VI) pre- interactions between orthophosphate and several condensed mo-
acidified to pH 1, in which orthophosphate displaces the central lybdates have been reported, each leading to the formation of a
labile molybdate unit from pre-existing dodecamolybdate to form different phosphomolybdate species [8,28]. It is reasonable to
11-MPA. This is clearly shown by the existence of an isosbestic point conclude, then, that this 'ground-up' mechanism proceeds via
between the two latter species [10]. 11-MPA then reacts with the initial reaction of orthophosphate with a low-Z molybdate species
displaced (and now acidified) molybdic acid unit to form 12-MPA; [10] to form a smaller phosphomolybdate, followed by rapid
the rate-limiting step is the final formation of 12-MPA from 11- equilibration to 12-MPA upon acidification via the condensation of
MPA. Mo(VI) (Fig. 3).

H2 O  2 PO3  3 Hþ


þ 4
12MoO2
4 þ 22H % ðMoO Þ
3 11 ðMoO4 Þ % ðMoO3 Þ11 ðPO4 Þ þ MoO2
4 % H3 PMo12 O40 (11)

The displacement mechanism (Eq. (11)) best matches the con- When 12-MPA is finally formed, it is inevitably protonated to
ditions normally used in MB methods, as the sample is typically some extent, contrary to early opinion [36]. Its pKa values in
introduced into a pre-mixed reagent solution in both batch-wise aqueous solutions have been determined as 2.4, 4.31 and 5.46 [37];
and flow-based methods. Furthermore, in MB methods, the ratio 12-MPA should therefore exist mainly as H3PMoO40 in MB reaction
of [Mo(VI)]/[P] in solution is high, thus accelerating the reaction. mixtures. At pH values where the trianion is prevalent, the complex
For example, a method that uses 10 mmol L1 Mo(VI) to determine decomposes in aqueous solutions [32].
1 mg L1 P (32 mmol L1) as orthophosphate will exhibit a [Mo(VI)]/
[P] ratio of over 300, whereas Murata and Ikeda in their Raman
2.3. Redox chemistry of PMB
spectroscopy work used only the stoichiometric ratio of 12 [10]. It
should be noted that if conditions change such that [Mo(VI)]/
2.3.1. Reduction of 12-MPA
[P]  14, 12-MPA will decompose into 11-MPA, 9-MPA and ulti-
The reduction of 12-MPA to form PMB is not a trivial process, as
mately [P2Mo5O23]6 [8,35]. It is therefore expected that a large
the nature of the product(s) is determined by a number of factors
excess of Mo(VI) over P will drive the phosphomolybdate equilibria
such as the reductant, pH, reaction time, temperature, and presence
to form 12-MPA almost exclusively, rather than leaving significant
of interfering ions. Knowledge of these factors is therefore of great
residual amounts of 11-MPA.
importance in developing MB methodology. In this discussion, the
The alternative mechanism, known as the 'ground-up' mecha-
degree of reduction of the PMB species is denoted using the
nism, occurs when acid is added to a pre-mixed solution of
number of electrons they are reduced by, i.e. PMB(2e) refers to the
orthophosphate and Mo(VI) and is much more rapid than the
two-electron reduced species, and PMB(4e) refers to the four-
displacement mechanism. This mechanism involves the assembly
electron reduced species.
of 12-MPA around an orthophosphate ion. Raman spectroscopy has
It is well-established that the redox behaviour of 12-MPA is
indicated that only MoO4 2 appears to exist in solution before
highly acid-dependent; the reduction of aqueous 12-MPA occurs in
acidification [10], yet it has previously been reported that ortho-
single electron steps at higher pH values, whilst this behaviour
phosphate binds so strongly to molybdate in solution that it cannot
changes to two-electron steps under acidic conditions [38e42] via
exist as the free ion in the presence of the latter [32]. Indeed,
proton-induced disproportionation of the odd-electron reduced
species [43]. Each two-electron reduction step of 12-MPA is
accompanied by the addition of two protons [38,39], and increasing
solution acidity shifts its reduction potential to more positive
values [38,39] since the overall negative charge of the complex is
decreased as it protonates. In fact, the correlation between the
anionic charge of 12-MPA and its first reduction potential is linear
[38,43]. Voltammetric analysis of 12-MPA in aqueous
0.2 mol L1 H2SO4 solutions (pH z 0.78) has shown that 12-MPA
may undergo two successive and reversible two-electron re-
ductions at þ0.31 V and þ0.17 V [42], which decrease to þ0.27 V
and þ0.13 V at pH 1 [44]. A third two-electron reduction at
about 0.08 V is possible but this renders the product highly un-
stable, resulting in immediate decomposition [42] (Fig. 5). It must
be noted that reduced heteropoly acids are unstable with respect to
re-oxidation by dissolved O2 [45]; their continued existence in
aqueous solutions depends on an excess of reductant to react
preferentially with O2, an aspect of reaction optimisation which is
discussed in Section 3.3.
Furthermore, the ability of Keggin-type structures to delocalise
Fig. 4. Influence of pH on the formation of phosphomolybdate species in an aqueous metal d-electrons influences their redox behaviour in an unusual
solution; [P] ¼ 1 mmol L1, [Mo(VI)] ¼ 12 mmol L1, l ¼ 420 nm. Whilst the absor-
bance above approximately pH 1 arises from a combination of phosphomolybdate
way; it is now well-established that each added electron pair is
species, note the sharp decline in absorbance towards lower pH values. Redrawn from delocalised across all twelve Mo atoms [46e48], thus avoiding
Ref. [32] with permission from Elsevier. multiple reduction of any individual Mo centre. In general, the
E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82 65

30000 detail, and generalisations have instead been made about the
stoichiometric behaviour of a given reductant, rather than the
Molar absorpƟvity (L mol-1 cm-1)

25000 actual MB species in solution [27,52]. The reaction conditions


required to form each different heteropoly blue, as well as their
20000 spectral properties, are discussed in the remainder of Section 2 and
are summarised in Fig. 6.
15000 The UVevisibleeNIR spectra of PMB species contain seven
individual absorption bands, each of which corresponds to a
10000 particular structural feature (Table 2), and the spectra of the
heteropoly molybdenum blue species (orthophosphate, arsenate
5000
and silicate species) all behave in a similar manner. However, only
two of these bands are of interest for MB methodology; namely,
0
the strong inter-valence charge transfer bands which arise due to
-0.1 0 0.1 0.2 0.3 0.4 0.5 0.6
electron exchange between Mo(V)and Mo(VI) centres. Contrary to
Applied potenƟal (V) some earlier claims, it is not the ded bands of Mo(V) which are
responsible for its intense absorption [53]; their nature as sym-
Fig. 5. Apparent molar absorptivity at 830 nm of a 12-MPA solution as a function of
applied potential during its electrochemical reduction [H2SO4] ¼ 0.2 mol L1, [12- metryeforbidden transitions marks them as orders of magnitude
MPA] ¼ 2 mmol L1. Adapted with permission from Tanaka et al. [42]. Copyright weaker than charge transfer bands, and they appear as little more
2015 American Chemical Society. than faint shoulders on typical MB spectra if they are discernible
at all (Table 2).
reduction of 12-MPA by n electrons (omitting further protonation) The molar absorptivities of the strong IVCT bands are signifi-
can be described by Eq. (12). cantly higher for PMB(4e) than PMB(2e) [54,55,57], but excessive
reduction yields PMB(6e) which immediately decomposes in acid
 n
H3 PMoðVIÞ12 O40 þ ne / H3 PMoðVIÞ12n MoðVÞn n ¼ 2; 4 [40,42,58]. This demonstrates that whilst a higher degree of
reduction is desirable for improving analytical sensitivity, over-
(12)
whelming reducing power should be avoided. The positions of the
intense IVCT absorption bands are dependent on a number of fac-
2.3.2. Spectral features of PMB tors discussed below.
It has frequently been observed that the use of different re- Molar absorptivity values of PMB in various reported methods
ductants e or even the same reductant under different conditions e have been used as a means of comparing analytical sensitivity be-
yields products with different UVevisible spectra and various tween methods. However, usage of this term as a comparative
apparent molar absorptivities, thus generating considerable measure is often incorrect. All methods which use non-metallic
confusion as to the ‘optimal’ reaction conditions and measurement reductants ultimately produce the same product with the same
wavelengths [40,49e51]. However, the nature of the absorbing innate molar absorptivity, [H4PMo12O40]3, or the 2e reduced
species produced by the reduction step is rarely described in any species as an intermediate when heating is not used (Fig. 6);

Non-metallic (α + β)-H3PMo12O40 Non-metallic


reductant '12-MPA' reductant, Sb(III)

β-[H2PMo12O40]3- 'PMB(2e-)'* [PSb2Mo12O40]- 'Sb2PMB(4e-)'


• ~ 700 nm (ε ≈ 7,000) • 880 nm (ε ≈ 20,000)
SnCl2 • 710 nm (ε ≈ 15,000)

Hea ng

β-[H3PMo12O40]4- 'PMB(4e-)' α-[PMo10Sn2O37]5- 'Sn2PMB(4e-)'


• 880 nm (ε ≈ 26,000) • 700 - 720 nm (ε ≈ 19,000)
• 710 nm (sh) • 620 nm (ε ≈ 16,000)

H+

β-[H4PMo12O40]3- 'PMB(4e-)' H+, Mo(VI)


H+ or
• 820 nm (ε ≈ 26,000)
• 660 nm (ε ≈ 10,000) (sh) hea ng

*Spontaneously
α-[PMo11SnOx]n- 'SnPMB(4e-)'
• 700 - 720 nm (ε ≈ 29,000) dispropor onates into 12-MPA
• 620 nm (sh) and PMB(4e-) under acidic
condi ons.

Fig. 6. Spectrophotometric properties of the various PMB species and the conditions under which they form. Molar absorptivities (ε) are in units of L mol1 cm1 (sh) ¼ absorption
shoulder.
66 E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82

Table 2
UV/visible spectral features of PMB(4e) species.

Wavelength (nm) Type of transition Assignment Band intensity

220 LMCT MoeO Strong, decreases with further reduction. Not specific to reduced species.
315 LMCT MoeOeMo bridges
550 ded Mo(V) Very weak, often obscured
760 ded Mo(V) Very weak, often obscured
600e700 IVCT Mo(V) / Mo(VI) Moderate, ε z 10,000 L mol1 cm1
700e900 IVCT Mo(V) / Mo(VI) Strong, ε z 26,000e34,000 L mol1 cm1
1500 IVCT Mo(V) / Mo(VI) Very weak, obscured

IVCT denotes an inter-valence charge transfer transition. LMCT denotes a ligand-metal charge transfer transition. Data obtained from Refs. [34,48,54e56].

differences in apparent absorptivity are simply due to the extent of the concentrations and proportions of reagents are altered
the reduction process at the time of measurement or the influence (Table 3). The question we seek to resolve here, then, is what these
of organic solvents [42,56,59]. The molar absorptivities of PMB absorbing species actually are and how their equilibria can be
species can be difficult to determine since quantitative formation of controlled in order to best optimise a MB method. These data are
a single PMB species is necessary. The difficulty lies in the nature of summarised in Fig. 6. Furthermore, since the spectral behaviour of
MB chemistry as one complex equilibrium system; rather than the various 12-heteropoly blues is very similar, studies on arseno-
stoichiometric reactions occurring, sufficient reagent must be and silico-molybdate species are useful for understanding their
introduced to perturb the equilibria to a desirable extent. Thus, if phosphorus counterparts.
one attempts to simply dissolve solid 12-MPA in acid and reduce it The positions of the two intensely absorbing IVCT bands are
without adding Mo(VI), very low apparent absorptivities will be determined by the degree of reduction, the protonation state of the
obtained [40], as most of the 12-MPA will in fact decompose in absorbing species, and whether or not a metallic reductant is used.
accordance with its formation equilibria (Fig. 3) [40,60]. Thus, The separation between the two bands is also dependent on these
sufficient Mo(VI) must be present to stabilise 12-MPA against hy- parameters, but more so on the degree of reduction; the IVCT bands
drolytic decomposition. of PMB(2e) are much more widely separated than those of
According to Tanaka et al. [42], electrochemically formed PMB(4e).
PMB(4e) exhibits an absorptivity of approximately As can be seen in Table 4, protonation has a profound effect on
25,000 L mol1 cm1 at 830 nm (Fig. 5). This study also indicates a the absorptivities and positions of the PMB(4e) IVCT bands. It is
molar absorptivity of ~11,000 L mol1 cm1 at potentials where now established that for four-electron reduced heteropoly acids,
PMB(2e) should be the main product. However, given that the first three protons are strongly acidic, whereas the remainder
PMB(2e) is expected to undergo acid-induced disproportionation are weakly acidic [58,61,87]. For example, the weak pKa values for
to 12-MPA and PMB(4e) [58,61] (Fig. 6), and that the UVevisible SiMB(4e) are 2.8, 3.8, 7.1 and 9.5 [58], whilst those for AsMB(4e)
spectrum of the product obtained at these potentials was identical are ~3, 4.5, 7.2 and 9.5 [61]. Given that orthophosphate MB methods
to that of PMB(4e), it can be concluded that the species observed are generally performed at pH < 1, it can be expected that PMB(4e)
here was not PMB(2e) at all, but PMB(4e) formed in approxi- will almost always be present as [H4PMo12O40]3, with the possi-
mately 50% yield. The molar absorptivity of PMB(2e) is likely to be bility of [H3PMo12O40]4 existing at very low acid concentrations.
around 7000 L mol1 cm1 at its absorption maximum of ~700 nm, This assignment is confirmed by the spectra of electrochemically
based on the ratio between peak intensities of similar heteropoly formed PMB(4e) [42], characterisation of pure PMB(4e) at pH 1
blue species [58]. [86], and ion-pair HPLC [56]. These two highly protonated forms of
The more Mo(VI) is added to the reaction, the more the for- PMB(4e) are amongst the most intensely absorbing of any PMB
mation of 12-MPA is favoured. The obvious question then is how species, with virtually identical absorptivities at their respective
much Mo(VI) can be added without the resulting gains in sensi- wavelength maxima (Fig. 6).
tivity being offset by direct Mo(VI) reduction, which is the phe- On the other hand, PMB(2e) is rarely desirable or even
nomenon responsible for the reagent blank in MB methods. In
keeping with the equilibria discussed at length above, the con-
centration of Mo(VI) needed to maintain constant Z is a non-linear
30 30000
function of acidity, whilst the acidity itself strongly affects the
Apparent molar absorpƟvity

stability of 12-MPA. At a given acid concentration, the amount of 25


12-MPA formation and reduction increases sharply with 26000
[Mo(VI)] (mol L-1)

(L mol-1 cm-1)

increasing Mo(VI) concentration, until a certain critical Mo(VI) 20


concentration is reached. At this point, the further formation of 22000
PMB tapers off and the direct reduction of Mo(VI) to form iso- 15
polymolybdenum blues begins [40]. Thus, it can be seen that 18000
10
whilst the Mo(VI) concentration can be easily optimised for each
acidity, sensitivity is quickly lost at higher acidities (Fig. 7) (i.e. Z 14000
5
values) due to 12-MPA decomposition into cations (Fig. 3). Note
that heating the reaction allows this effect on 12-MPA formation 0 10000
to be overcome to a great extent (Table 3), but at the expense of 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
sampling rate. [H+] (mol L-1)

Fig. 7. Plot of acidity against the maximum tolerable Mo(VI) concentration (C) before
direct reduction of Mo(VI) to isopoly MB begins, and the apparent molar absorptivity of
2.3.3. Nature of the reduced products the SnPMB(4e) product at 700 nm ( ) under these conditions (Data obtained from El-
It is well-known that the UVevisible spectrum of a particular Shamy and Iskander [40]. [12-MPA] ¼ 40 mmol L1 for molar absorptivity data,
method's PMB product(s) is modified, potentially dramatically, if [SnCl2] ¼ 880 mmol L1).
Table 3
Comparison of chemical and spectrophotometric parameters of batch MB methods.

Authors (year) Acid [Acid] [Mo(VI)] [Reductant] Temperature ( C), lmax ε Notes Ref
(mmol L1) (mmol L1) (mmol L1) (reaction time) (nm) (L mol1 cm1)

Tanaka (1982) H2SO4 200 2 (12-MPA) Electroreduction e 830 25,000 Optical path length uncertain [42]
Fontaine (1942) H2SO4 1000 36 SnCl2: 210 100 (20 min) 820 28,000 [62]
Sims (1961) HCl 650 17 SnCl2: 2.43 e 815 26,400 [63,64]
El Sayed (2001) H2SO4 50 1.3 SnCl2: 1.04 e 700 23,000 Broad peak, shoulder [65]
at ~ 820 nm
Kriss (1971) H2SO4 300 5 SnCl2: 5 (<3 h) 680 20,400 [66]
Levine (1955) HCl 600 11 SnCl2: 0.16 e 735 19,000 [67]
El-Shamy (1973) HCl 840 20 & 0.08 SnCl2: 1.93 e 810 17,000 Peaks in equilibrium via [40]
(12-MPA) isosbestic point at 780 nm
HCl 360 20 & 0.08 SnCl2: 1.93 e 720 16,000 [40]
(12-MPA)
HCl 450 0.165 SnCl2: 0.66 e 810 4400 Broad shoulder around 720 nm [40]
(12-MPA)
HCl 280 0.24 SnCl2: 1.44 e 700 1200 Shoulder at 820 nm [40]
(12-MPA)
Hesse (1968) H2SO4 300 31 SnCl2: 0.13 (20 min) 730 25,500 [68]
HS: 2.31
Drummond H2SO4 127 2.4 AA: 9.6 (12 h) 820 26,760 [29]
(1995)
Han (1995) H2SO4, 330, 14 AA: 54 45 (20 min) 825 26,000 [69]
HCl 170
Chen (1956) H2SO4 300 14 AA: 57 37 (90 min) 820 25,000 [70]
Lowry (1946) pH 4 5.7 AA: 5.7 (5 min) 860 4600 Very similar absorbance from [71]
700 to 900 nm
Katewa (2003) H2SO4 500 14 AA: 11.4 (60 min) 820 26,100 [49]
HS: 15.4
Sims (1961) H2SO4 500 10.3 HS: 0.46 100 (unspecified) 815 32,300 [63]
Ganesh (2012) H2SO4 500 39 HS: 9.61 60 (30 min) 830 29,000 [72]
Boltz (1947) H2SO4 500 10.3 HS: 0.46 93 (10 min) 830 26,400 [73]
Huey (1967) HClO4 1200 24 HS: 1.8 100 (15 min) 805 14,000 [60]
Burton (1956) H2SO4 30 2.95 Metol: 2.9 100 (2 h) 860 29,900 [50]
Na2SO3: 87
H2SO4 200 2.95 Metol: 2.9 100 (2 h) 820 29,300
Na2SO3: 87
H2SO4 30 2.95 Metol: 1.5 (2 h) 720, 830 (sh) 3900 Extremely broad absorbance
Na2SO3: 44 peak
Sims (1961) HClO4 920 143 Metol: 0.33 100 (10 min) 820 18,500 [63]
NaHSO3: 30
Na2SO3: 4.0
Harris (1954) HClO4 480 11.3 Metol: 1.16 e 820 3900 [63,74]
NaHSO3: 96
Na2SO3:15
El Sayed (2001) H2SO4 30 11 ANS: 0.11 90 (30 min) 830 32,000 [65]
Sims (1961) HClO4 920 143 ANS: 0.33 100 (10 min) 820 28,100 [63]
NaHSO3: 30
Na2SO3: 4.0
Griswold (1951) H2SO4 500 14.2 ANS: 0.42 100 (10 min) 820 27,000 [75]
NaHSO3: 570
Na2SO3: 16
Fiske (1925) H2SO4 250 14 ANS: 0.42 e 730 3900 [76,77]
NaHSO3: 57.7
Sims (1961) HClO4 920 143 DAPH: 0.33 100 (10 min) 820 34,500 [63]
NaHSO3: 30
Na2SO3: 4.0
Huo (2012) H2SO4 300 10.3 Thiamazole: 0.245 e 710 1000 [78]
Kriss (1971) H2SO4 300 5 Fe(II): 5 (<3 h) 700 12,800 [66]
Kriss (1971) H2SO4 30 5 HQ: 5 (<3 h) 700 12,800 [66]
Salem (1991) H2SO4 76 2.8 Oxalyldihydrazide: 100 (10 min) 880 33,000 Shoulder at 820 nm [79]
0.017
Gupta (1981) H2SO4 200 5.4 AA: 4.8 (10 min) 882 25,670 [80]
Sb(III): 0.066
Going (1974) H2SO4 150 3.5 AA: 4.5 (10 min) 880 22,400 ε ¼ 26,400 (840 nm) without Sb. [81]
Sb(III): 0.8
Pai (1990) H2SO4 200 5.4 AA: 4.8 (30 min) 880 22,400 [82]
Sb(III): 0.066
Harwood (1969) H2SO4 200 2.7 AA: 23 (10 min) 890 20,600 [83]
Sb(III): 0.32
Edwards (1965) H2SO4 57 1.5 AA: 1.4 (10 min) 880 20,400 About 4% EtOH required to [84]
Sb(III): 0.15 dissolve Sb-PMB precipitate
for P > 1 mg L1
Murphy (1962) H2SO4 200 5.4 AA: 4.8 (10 min) 882 22,400 [85]
Sb(III): 0.066
Drummond H2SO4 127 2.4 AA: 9.6 (1.5 min) 880 21,680 [29]
(1995) Sb(III): 0.058

Table abbreviations: AA; ascorbic acid, HS; hydrazinium sulfate, HQ; hydroquinone, ANS; 1-amino-4-naphthol-2-sulfonic acid, DAPH; 2,4-diaminophenol dihydrochloride, 12-
MPA; 12-molybdophosphoric acid, EtOH; ethanol, Metol; 4-(methylamino)phenol sulfate, PMB; phosphomolybdenum blue.
68 E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82

Table 4
Effect of protonation on IVCT absorption maxima and approximate molar absorptivities for b-[XnþMo12O40](12n). Roman numerals denote the number of electrons the
compound is reduced by.

Protonation and Absorption maxima (nm) in aqueous solution. Approximate molar absorptivities (L mol1 cm1) and peak shapes are given in
reduction state parentheses where possible; (br) ¼ broad absorption peak, (sh) ¼ absorption shoulder.

X ¼ As5þ X ¼ Si4þ X ¼ P5þ


IV 1110 (6600), 741(8400) 1050, 765
HIV 990 (9600), 752 (9000) 1000 (br), 760 (br)
H2IV 935 (16,000), 719 (sh) 925, 720 (sh)
H3IV 885 (25,000), 704 (sh) 880, 700 (sh) 880 (26,000), 700e710 (sh)
H4IV 840 (25,000), 667 (sh) 820, 660e680 (sh) 820 (26,000), 660 (sh)
II 1120 (br), 720 1030, 760
H2II 1050, 680 ~1000, 730 (7000, br)

Data obtained from Refs. [42,50,54e56,58,61,76,79,86].

observed; it has considerably lower molar absorptivities at its distinctly different conditions [90e92], but has been overlooked
peak wavelengths compared with those of PMB(4e) which ren- in analytical methods based on phosphomolybdate. This is prob-
ders it an inferior chromophore, and its oxidising ability will ably because b-MPA is less stable than its silicon counterpart [88],
readily continue the reduction to PMB(4e) in the presence of and whilst evidence suggests that the two isomers form simul-
further reductant. PMB(2e), like its As and Si analogues, can be taneously from different molybdate precursors in fresh 12-MPA
presumed to also be unstable with respect to acid-induced solutions [89,91,93], b-MPA isomerises to a-MPA within minutes
disproportionation, spontaneously forming PMB(4e) and 12- [89]. However, isomerisation from a / b appears to begin in
MPA [58,61]. Therefore, this species is generally only obtained earnest once the reduction process begins, and voltammetric data
via electrochemical means in organic solvents [54] or transiently for aqueous 12-MPA indicate that nearly 10 min after the first
with a kinetically slow reductant [74,77]; on the rare occasions reduction of pure a-MPA, the ratio of a- and b-isomers is nearly
when it is encountered, its IVCT bands are distinctively separated 1:1 [42]. Because the b-isomer is the thermodynamically favoured
by 350e400 nm, such that only the broad ~ 700 nm band is form of the reduced species, heating will accelerate the a / b
observed in the visible region [54,56]. transformation of PMB [94]. It should be noted, however, that the
isomerism of PMB applies strictly only to [HnPMo12O40](7n),
2.3.4. Isomerism of 12-MPA and its reduced forms which is generally only obtained from the heating of organic re-
12-MPA is typically encountered as one of two isomers; the a- ductants. Metallic reductants are capable of altering the compo-
isomer is the nominal Keggin structure and is the stable form of sition of PMB and locking it into one particular isomer, a
unreduced 12-MPA, whereas the b-isomer, obtained by a 60 phenomenon discussed below.
rotation of a Mo3O13 group, is the more stable form of the reduced
species (Fig. 8) [88,89]. The optical and redox properties of the two 2.3.5. Organic reductants
isomers differ, in that the b-isomer is reduced at more positive MB methods which use only organic reductants or hydrazine
potentials and exhibits more intense IVCT bands [34,55,89]. This sulfate (the ‘non-metallic’ reductants) in combination with heating
phenomenon is well-known in silicomolybdate chemistry where demonstrate a clear absorption maximum at 820 nm (Table 3), with
the two isomers can be isolated in aqueous solution under a discernible shoulder at 660 nm. Electrochemical reduction of 12-

Fig. 8. Polyhedral representations of the a- and b-isomers of the Keggin structure. Point group symmetries are indicated, and the labels in parentheses distinguish between the
three types of O atoms (terminal, bridging, or part of the central tetrahedron). The b-isomer is obtained from the a-isomer by a 60 rotation of one Mo3O13 subgroup. Reprinted with
permission from Ref. [88]. Copyright 2015 American Chemical Society.
E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82 69

MPA also demonstrates these features [42], which are attributed to [97]. The Sn(IV) substitution process is not observed with 12-MPA
the species b-[H4PMo12O40]3 (Table 4). An exception to this is seen before reduction, nor does Sn(II) substitute for Mo after reduction.
in the method of Salem [79] where the unusually low acid con- Thus, upon four-electron reduction of 12-MPA by SnCl2, the
centration encourages the predominance of b-[H3PMo12O40]4 immediate product appears to be a-[PMo10Sn2O37]5 [94,96],
instead, with an absorption maximum at 880 nm instead of the denoted as Sn2PMB(4e), with an absorption peak between 700
usual 820 nm. and 720 nm (ε z 19,000 L mol1 cm1) and a pronounced shoulder
A heating step is necessary with these reductants due to the at 620 nm.
slow kinetics of 12-MPA reduction when using them, possibly
because of the complex reaction steps required for their oxidation a  H3 PMo12 O40 þ 4e þ 2Sn4þ þ 3H2 O
[27] or, more likely, their weak reducing power in acidic solutions
(Eq. (13)) [95]. /a  ½PMo10 Sn2 O37 5 þ 2MoO3 þ 9Hþ Eh z þ 0:7 V
(14)
Dehydroascorbic acid þ 2Hþ þ 2e $Ascorbic acid
(13) The rapid kinetics of this reduction process are attributed to a
Eh ¼ þ0:33VðpH 1Þ
combined reduction/substitution step with a much more positive
It should be pointed out that whilst virtually all methods which reduction potential than for the direct reduction of 12-MPA alone
use these reductants exhibit molar absorptivities of at least (Eq. (14)) [97]. Over time, this species spontaneously undergoes the
25,000 L mol1 cm1, several of them reportedly exhibit absorp- loss of one Sn(IV) ion to form the incompletely characterised
tivities of 32,000e35,000 L mol1 cm1 (Table 3), the highest such SnPMB(4e) species, as has been observed with its silicate analogue
values for PMB reported in the analytical chemistry literature. [94]. SnPMB(4e) absorbs more intensely at 720 nm than
However, there exists considerable variation in the molar absorp- Sn2PMB(4e) whilst the absorptivity of the 620 nm band hardly
tivity data reported in the literature, which is of some concern. For changes, becoming a barely noticeable shoulder (Fig. 9) [96], and
example, two studies using the Murphy and Riley method [80,85] this mono-tin species is the main reduction product monitored in
obtained apparent molar absorptivities differing by methods using SnCl2 reduction. The spectral features of both
3000 L mol1 cm1 (Table 3), whilst the hydrazine sulfate methods products are reported in the analytical literature, i.e. Sn2PMB(4e)
of Sims and Boltz [63,73] reported absorptivities [85,100] and SnPMB(4e) [66,99]. The nature of the transformation
6000 L mol1 cm1 apart despite using very similar conditions and of Sn2PMB(4e) to the more intensely absorbing SnPMB(4e) re-
wavelengths which should have yielded very similar absorptivities mains unclear, although it is not a redox process since Sn2PMB(4e)
(Table 3). These significant differences are noteworthy, since there should be stable with respect to oxidation [96]. However, this
is of course only one actual 'molar absorptivity' for a particular process is spontaneous, and is likely the result of attack by acidified
compound at a given wavelength. The differences in reported Mo(VI) species [97].
absorptivitives are attributed to variations in handling procedures, SnPMB(4e) may ultimately be hydrolysed to [H4PMo12O40]3
extent of reaction, uncorrected blank formation, the spectropho- over time, although this process is accelerated with either heating
tometric instrumentation used, and assumptions about apparent [62,94] or further acidification [40], indicated by the emergence of
absorptivity versus actual molar absorptivity. the 820 nm peak (Fig. 6). This process therefore appears to involve
Aside from methods using SnCl2 as a reductant (discussed the displacement of Sn(IV) by molybdic acid, which seems to occur
below), those which report an absorbance maximum at ca. 720 nm in the conversion of Sn2PMB(4e) to SnPMB(4e) as well. These
show vastly inferior molar absorptivities (Table 3). These methods reactions are summarised in Fig. 6.
do not use heating; as a result, the predominant reduction product It should be pointed out that SnCl2 only undergoes the chem-
is the two-electron reduced species [H2PMo12O40]3, and the actual istry discussed above when it reduces a-isomers of heteropoly
extent of 12-MPA / PMB(2e) reduction is expected to be quite acids, and all of the stannomolybdate species discussed above are
low as well. Furthermore, this species is only transient, and will themselves a-isomers [96]; the reduction of b-isomers yields
exhibit an unstable UVevisible spectrum over time as it dispro- PMB(4e) directly without incorporation of Sn(IV). Given the
portionates to PMB(4e) and 12-MPA.

30000
2.3.6. Metallic reductants SnCl2
Molar absorpƟvity (L mol-1 cm-1)

The use of metallic reductants has a significant effect on MB 25000


chemistry, as the metals themselves may be incorporated into the Ascorbic acid
reduced product. Sn(II) and Sb(III) are the main species used in this 20000
Ascorbic acid / Sb(III)
regard; the former reduces 12-MPA very rapidly, whereas the latter
is used to accelerate the reduction by an otherwise kinetically slow 15000
reductant such as ascorbic acid.
SnCl2 is very widely used due to its fast reduction kinetics which 10000
obviates the need for heating. However, it is quite clear that 12-MPA
reduction with SnCl2 yields a distinctly different visible spectrum to 5000
that of [H4PMo12O40]3 (Fig. 9) which cannot be accounted for via
(de)protonation. In fact, instead of the conventional four-electron 0
reduction of 12-MPA, the presence of newly-formed Sn(IV) ions 350 400 450 500 550 600 650 700 750 800 850 900
permits a different and much more favourable reaction in which Wavelength (nm)
the four-electron reduced product, still in the a-form, immediately
reacts with two Sn(IV) ions which substitute for Mo(VI) [94,96]; Fig. 9. Comparison of product spectra for the three most commonly used reductants in
whilst this behaviour has only been studied on 12-MSA, it appears MB methods. Ascorbic acid/Sb(III); prepared according to the procedure described in
Standard Methods for the Examination of Water and Wastewater [98]. Ascorbic acid;
analogous to that of 12-MPA. The vastly more positive reduction prepared as for ascorbic acid/Sb(III) procedure without the addition of Sb(III) and with
potential of 12-MPA in the presence of Sn(IV) has been discussed as 
heating at 100 C for 10 min SnCl2; prepared using the working concentrations re-
the likely cause of the fast reduction kinetics observed with SnCl2 ported in Ref. [99]. All solutions contained 500 mg L1 P as orthophosphate.
70 E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82

tendency for a- and b-MPA to form concurrently from orthophos- complex is probably [PSb2Mo10Ox]n, wherein Sb(III) is substituted
phate and acidified molybdate [89], and the equilibrium between for Mo(VI), this conclusion appears to be based only on the findings
SnPMB(4e) and PMB(4e), it is to be expected that methods using of Going and Eisenreich in 1974 [81], which were derived from an
SnCl2 will form a mixture of PMB(4e), SnPMB(4e) and average measured Mo:P ratio of 11.4:1 and the observed stoichi-
Sn2PMB(4e) giving absorbing species with wavelength maxima ometry of other heteropoly species. A recent ESI-MS study
100 nm apart. This effect is clearly shown in the spectra recorded by addressing this question found no evidence of [PSb2Mo10O40]n or
a number of authors [40,65,85]. If an undesirable shoulder is seen at protonated derivatives under any conditions, but did observe peaks
820 nm, or the absorption peak is unusually broad, the acidity in aqueous solutions of pH z 0.8 corresponding to
should be decreased to favour the formation of only the Sn- [PSb2Mo12O40]n and species with one or both antimony atoms
substituted products [40]. removed in organic extracts of this solution [106]. Whilst these
Antimony (Sb) has been of great importance in MB method- results have not yet been addressed by other studies, they none-
ology since the work of Murphy and Riley in 1962 [85]. The theless suggest that Sb does not replace Mo in the Keggin structure
presence of Sb(III) was originally found to greatly accelerate the as originally thought; the twelve Mo atoms remain regardless of Sb
reduction of 12-MPA by ascorbic acid, with the reaction going to addition.
completion in approximately 10 min without the need for heating Going and Eisenreich originally suggested that the dramatic
[85,101] and yielding a product which remains stable for hours. increase in the reduction rate of 12-MPA in the presence of Sb(III)
With sufficient ascorbic acid, this time can even be decreased to may be due to either a structural change in the heteropoly acid
1.5 min [29]. The development of methods using Sb(III) was of which facilitates reduction, or to Sb(III) acting as an electron relay
great practical significance, as it enabled the comparatively rapid by first reducing 12-MPA and forming Sb(V) which is then imme-
determination of orthophosphate without lengthy heating steps. diately reduced by ascorbic acid to Sb(III) [81]. An electron relay
Whilst other organic reductants such as hydroxylamine salts [102] mechanism would require Sb(III) to act as a reductant before being
behave in the same manner as ascorbic acid in the presence of subsequently reduced by ascorbic acid; Sb(III) is capable of
Sb(III), ascorbic acid's popularity as a reductant predates even reducing 12-MPA directly but only at high temperatures
Murphy and Riley's work by decades [103], presumably due to its [81,107,108], and no reaction is observed under the conditions ex-
ready availability and non-hazardous nature. Just as for the case of pected for the MB reaction [106]. This suggests, therefore, that
SnCl2, the presence of Sb(III) has a significant effect on the visible Sb(III) does not act as an electron relay.
IVCT bands of the PMB product, countering the early view that Sb(III) does not interact with acidified molybdate, nor does it
Sb(III) served only as a catalyst. 12-MPA reduced in the presence of alter the IVCT bands of pre-formed PMB [81]; it does not even
Sb(III) exhibits an absorption maximum at ca. 880 nm, with a appear to form an adduct with unreduced 12-MPA at ambient
distinct second peak at 710 nm and a shoulder in the vicinity of temperature [106], contrary to earlier assumptions [109]. However,
550 nm (Fig. 9) [80,85,104]. These features are similar to the it does redshift the characteristic UV charge transfer bands of 12-
[H3PMo12O40]4 species (Table 4), except that the 710 nm band in MPA, as do a number of other metals [81], suggesting at least
the Sb(III) product is a distinct peak. weak coordination to the oxygen atoms of 12-MPA. The full
It has long been known that antimony becomes part of the Sb2PMB(4e) adduct seems to form only during 12-MPA reduction
reduced complex and a central element in the stoichiometry of [106]. Heteropoly complexes which contain metals as counter-
the reaction. However, the original 1:1 Sb:P ratio determined by cations are known to exist as chargeetransfer complexes, in
Murphy and Riley [85] has been called into question by more which the metal acts as the electron donor and 12-MPA is the
recent and reliable results that show a 2:1 ratio of Sb:P in the electron acceptor; isolation of antimony salts in the solid phase has
reduced species [80,81,105] instead. This conclusion is in excellent shown that Sb(III) can indeed partially reduce 12-MPA when in this
agreement with observations about method linearity; a compar- configuration, although again only after a heating step [107,110].
ison of Sb(III) MB methods has shown that the linear range is However, the observation that Sb(III) does not bind to PMB after
limited by antimony concentration, and has a cut-off exactly reduction would seem to discount it acting only as a countercation
where the Sb:P ratio becomes less than 2:1 [83]. This phenome- or loosely coordinating species. Taken together, all of the above
non has been exploited to extend the linear range to more than evidence suggests that much like in the case of SnCl2, Sb(III)-
three times that of Murphy and Riley's method [83], although facilitated reduction is a reaction distinct from the normal four-
adding excessive amounts of Sb(III) results in solubility problems, electron reduction of 12-MPA. Interestingly, in the presence of
presumably due to hydroxide salts [83,85] or Sb(III) - Mo(VI) salts both Sb(III) and Sn(II), only the SnPMB(4e) product is observed,
(discussed in Section 3.4). It should be noted that this linear range probably due to the much faster reduction process involving sub-
is only with respect to absorbance at 880 nm; the further addition stitution of Sn(IV) into the product structure [111]. Despite these
of orthophosphate results in the additional formation of the observations, the precise mechanism of Sb2PMB(4e) formation
familiar [H4PMo12O40]3 species with IVCT bands at 820 and and the redox potential for the Sb(III)-facilitated reduction remain
680 nm, giving the false impression that the 880 nm band blue- unknown.
shifts with added orthophosphate due to the rising 820 nm Heating is sometimes applied in methods using Sb(III) to further
band [51,81,83]. enhance the rate of reduction [82,112]. However, heating of this
The structure of the reduced phosphoantimonylmolybdate reaction can be problematic. Not only does it increase the impact of
species Sb2PMB(4e), and the cause of its accelerating effect on the silicate interference [82,109] as would be expected of any method
reduction kinetics, are not clearly established. The molar absorp- involving heating (see Section 4.1.2), but higher reaction tempera-
tivity of Sb2PMB(4e) at 880 nm compared with the conventional tures also appear to degrade the Sb2PMB(4e) product [109,112]. If
[H4PMo12O40]3 product at 820 nm under the same conditions is the product is already fully formed and is exposed to successively
actually lower by 20e30% [29,81,85]; this trend is clearly confirmed higher temperatures, the UVevisible spectrum of the product ex-
in Table 3 and in Fig. 9. Nevertheless, molar absorptivities in the hibits lower apparent molar absorptivities at 710 and 880 nm with
vicinity of 20,000 L mol1 cm1 still suggest a four-electron an isosbestic point at 935 nm [112]. This evidence suggests that at
reduced product. Aside from the Sb:P ratio of 2:1, the stoichiom- high temperatures, the Sb2PMB(4e) product equilibrates with one
etry of the entire Sb2PMB(4e) complex has never been established or more species with an absorption maximum at wavelengths
with certainty. Whilst it has been held for some time that the longer than 950 nm (see Section 2.3.2). However, this spectral
E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82 71

change was found to be almost entirely reversible upon cooling the invariant across all solutions of the same [Hþ]/[Mo(VI)] ratio or one
reaction mixture again. In contrast, if the reduction itself is per- of the two variables should not affect the system at all [114].
formed at elevated temperatures, the concentration of Truesdale and Smith clearly showed that the pH varies considerably
Sb2PMB(4e) appears to decrease irreversibly, with considerable between solutions of various compositions with the same [Hþ]/
instability in absorbance at 880 nm during the reaction [112]. Pai [Mo(VI)] ratio [114], and it is self-evident from any study optimising
et al. recommend that MB methods using Sb(III) should not exceed the MB method that both parameters strongly affect the system. In
a reaction temperature of 35  C. Strickland's work, however, the narrow range of Mo(VI) concen-
trations used did not affect product formation significantly, and the
3. MB method optimisation [Hþ]/[Mo(VI)] ratio was thus effectively developed as an alternative
acidity scale within a strict range of Mo(VI) concentrations
The parameters of the MB reaction have traditionally been [114,115]. In fact, Truesdale and Smith commented that since the
optimised for batch methods according to the following goals: [Hþ]/[Mo(VI)] ratio is an inappropriate fundamental framework for
describing MB chemistry, its adoption has served to mislead further
 Maximal sensitivity to orthophosphate investigation of the system [91]. The use of this ratio has also been
 Minimal blank absorbance inconsistent between methods using monoprotic and diprotic
 Broad linear calibration range acids; older studies using the normality scale for acid concentration
 Stability of colour (i.e. absorbance) during the measurement clearly assume that both protons of H2SO4 fully dissociate in the
interval reagent solution, when in fact the second pKa of H2SO4 is only 1.99
 Fast PMB formation [116].
No single optimised combination of Hþ and Mo(VI) concentra-
However, the goals of flow-based techniques such as flow in- tions is clearly evident in the literature. Methods involving heating
jection analysis (FIA), reverse flow injection analysis (rFIA) and tend to dramatically reduce the influence of reagent concentrations
sequential injection analysis (SIA) differ considerably from those of on the apparent absorptivity due to faster reaction kinetics. SnCl2
batch, automated discrete analyser (DA) and segmented continuous methods raise the question of a mixture of absorbing products and
flow analyser (SFA) methodology [113]. Whilst the demands for the reported absorptivities for Sb(III) methods using the same re-
sensitivity, low blank absorbance and a broad linear range remain, action conditions are inconsistent (Table 3). It is therefore of much
the temporal stability of the product is no longer of particular greater practical use to recommend a procedure for the optimisa-
importance since sample detection occurs at a fixed point during tion of the MB method based on the principles discussed in this
the reaction, i.e. under non-equilibrium conditions. However, re- review.
action kinetics are still important in flow systems; whilst the extent Whilst heating the MB reaction typically renders the yield of
of reaction is inherently reproducible in these systems, it is obvi- PMB rather insensitive to the acid and Mo concentrations used
ously advantageous to maximise the extent of reaction in a short (Table 3), it is more practical to perform the reaction at room
timeframe to enhance sensitivity. It is therefore unsurprising that temperature. Furthermore, prolonged heating appears to decom-
the SnCl2 and Sb(III) methods now dominate the literature on both pose both PMB(4e) and Sb2PMB(4e) [60,112], and any heating at
flow-based and batch orthophosphate detection using MB, as the all accelerates the spontaneous hydrolysis of the SnPMB(4e)
reduction reaction occurs almost instantaneously with the former product [62,94] obtained from methods using SnCl2 as the reduc-
reagent, and within a few minutes with the latter. Furthermore, tant. Without heating, however, the effects of reagent concentra-
whilst the zero dispersion of SFA and DA methods allows for the tions on reaction kinetics and equilibrium positions become more
exact replication of optimised solution conditions from batch important. At this point, it is important to highlight the trend
methods, the substantial dispersion encountered using FIA, rFIA or shown in Fig. 7; a greater Mo concentration will always require a
SIA makes the same task impossible using these methods. Rather greater acidity to mitigate the reagent blank (discussed in Section
than dealing with discrete volumes of reaction mixture, such sys- 3.2), but the loss of sensitivity due to 12-MPA decomposition at
tems inherently create concentration gradients between neigh- these higher acidities is significant. Thus, whilst the acidity can
bouring liquid zones, which are further subjected to a parabolic easily be optimised for any given Mo concentration, method
flow profile arising from friction with the walls of the manifold sensitivity will necessarily decrease as [Mo(VI)] is increased.
tubing. Therefore, the concentration of Mo(VI) must be sufficiently high to
stabilise 12-MPA at P concentrations appropriate for the analysis,
3.1. Reagent concentrations but not so high as to warrant deleterious concentrations of acid.
Going and Eisenreich originally recommended that Mo(VI) be used
The optimisation and re-optimisation of MB methods has in at least a 40-fold molar excess over the maximum P concentra-
largely been based on the widespread and erroneous assumption tion expected [81] although their experiments focused on the [Hþ]/
that MB chemistry can be adequately described by the ratio of [Hþ]/ [Mo(VI)] ratio at different Mo(VI) concentrations. Nevertheless, this
[Mo(VI)], at which the behaviour of the reaction supposedly re- recommendation broadly agrees with other studies which have
mains constant regardless of the actual reagent concentrations found that Mo(VI) concentrations as low as 3e5 mmol L1 are
used. The use of this ratio appears to have originated in the work of sufficient for extensive 12-MPA formation from approximately
Strickland [114,115], and has been used ever since as an attempt to 1 mg L1 P (Table 3). The lowest possible acidity should be used,
summarise the optimal conditions for both the formation and such that PMB decomposition is minimised but direct Mo(VI)
reduction of 12-MPA, as well as blank minimisation. However, this reduction is still prevented or mitigated. El-Shamy and Iskander
ratio does not describe 12-MPA formation in solution due to the found that at 5 mmol L1 Mo(VI), an acidity of only 0.20 mol L1 HCl
speciation of Mo(VI) and the non-linear variation of Z with [Hþ] and was effective in preventing isopoly MB formation [40]. However,
[Mo(VI)] discussed in Section 2. Pai et al. [82] have clearly whilst higher acidities at this Mo concentration drastically inhibi-
demonstrated this by showing that PMB formation is not consistent ted PMB formation, no acidities lower than this value were
in solutions where [Hþ] and [Mo(VI)] are varied but [Hþ]/[Mo(VI)] examined.
remains constant. If the [Hþ]/[Mo(VI)] ratio were an accurate means Reductant concentrations vary widely across the literature
of describing molybdate chemistry, either the pH should be (Table 3) depending on the species chosen, and typically tolerate
72 E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82

much more variation without deleterious consequences reaction times, blank minimisation, avoidance of heating) are
compared with the acid and molybdate concentrations. For effective in mitigating isopoly MB formation. Coating problems are
example, ascorbic acid has traditionally been used in large excess typically due to the limited solubility of PMB species instead, which
[81] for kinetic reasons; there is evidence to suggest that the is discussed in more detail in Section 5.
reduction of 12-MPA by ascorbic acid is first order in the latter Just as for the reduction of 12-MPA, the reduction of Mo(VI)
[27], and given its weak reducing ability in acidic solutions [95], a depends strongly on the acidity, Mo concentration and reductant
high concentration of ascorbic acid is indeed expected to be of concentration. As a general rule, 12-MPA can be effectively reduced
some practical benefit. Whilst this approach is effective in Sb(III) under conditions where Mo(VI) alone cannot; isopoly MB species
methods [29] where complete 12-MPA reduction within a few are easily formed between pH 1.0e2.5 [3,118], whereas the for-
minutes is desirable, it is probably unjustified when heating is mation and reduction of 12-MPA can tolerate higher acidities and
used. In MB methods involving heating, sub-millimolar concen- much lower Mo concentrations. Isopoly MB formation is favoured
trations of an organic reductant are sufficient for extensive PMB with higher Mo concentrations, lower acidity, and higher reductant
formation in a similar timeframe to the Sb(III) methods (Table 3). concentrations; MB method optimisation is thus a compromise
Even at room temperature, SnCl2 or Sb(III) methods require only between maximising 12-MPA formation and reduction and mini-
low millimolar concentrations of reductant for effective use; most mising isopoly MB formation (Fig. 7).
SnCl2 studies report reductant concentrations of It is also interesting to note that in methods where a combined
1.0e2.5 mmol L1 SnCl2, and Sb(III)/ascorbic acid methods are reagent solution containing acid, molybdate and reductant is pre-
effective with around 5 mmol L1 ascorbic acid. Even a 1 mmol L1 pared, such as those reported in Standard Methods for the Exam-
reductant concentration represents an approximately 30-fold ination of Water and Wastewater [98], the reagent solution tends to
molar excess over 1 mg L1 P. Clearly, the question is not one of exhibit a yellow colour [91] due to absorption bands at 490 and
stoichiometry, but of reduction potential and kinetics. Excessive 385 nm. These bands can also be observed in some aged PMB
reductant concentrations have long been known to induce direct products, manifesting as a green hue, and reflect the ded transi-
Mo(VI) reduction after 12-MPA reduction is apparently complete tions of dimeric Mo(V) present as [Mo2O4(OH2)6]2þ [121,122]. The
[67,73,117], and to ‘over-reduce’ 12-MPA to colourless decompo- presence of this species indicates the direct reduction of Mo(VI)
sition products [40,42,79]. under conditions where isopoly species cannot form, due to either
the higher acidity of a combined reagent solution or the low Mo(VI)
3.2. The reagent blank: isopoly molybdenum blue species concentration used in some methods. Mo(V) halide complexes with
a similar yellow colour have also been reported to form in marine
Care must be taken in any MB method to minimise the reagent sample matrices [123]. Whether or not the [Mo2O4]2þ species itself
blank absorbance, which arises from the direct reduction of Mo(VI). acts as a reductant or is directly involved in forming PMB is
In contrast to the comparatively small Keggin structure, reduction currently unknown, but it does not interfere with the spectro-
of Mo(VI) alone yields a variety of giant isopolymolybdates with photometric determination of P. 12-MPA itself is also attributed a
proportions of Mo(V) varying from 15 to 20% [118], such as the 'Big faint yellow colour, although this appears to arise from the tail of
Wheel' (HxMo154On462) and the 'Blue Lemon' the 315 nm LMCT band extending into the visible region in
[HxMo368O1032(H2O)240(SO4)48](x48). The latter species forms concentrated 12-MPA solutions.
only in the presence of sulfate ions, and is likely to be a major
contributor to many MB reagent blanks given the widespread use of 3.3. Product stability
sulfuric acid in MB methods. These isopoly species are interferents
for two reasons. Firstly, since they possess some of the same The stability of the colour in MB chemistry is attributable to three
structural motifs as reduced 12-MPA, they also possess intense IVCT factors; the extent of reduction at the time a measurement is taken,
absorption bands in the same region of the visible-NIR spectrum, the stability of the product and the availability of excess reductant to
generally around 750e780 nm and 1050e1100 nm [3,118,119]. protect PMB from re-oxidation by dissolved O2. For example, it is
Wheel-type structures have a comparatively well-defined peak in frequently noted in studies which make use of organic reductants
this region, whereas other isopoly MB species have exceptionally without heating that the absorbance of the product continually in-
broad and ill-defined absorption bands. Since the molar absorp- creases for periods up to several hours [49,63,70,74]; given the slow
tivity of MB species scales linearly with the number of Mo(V) reduction kinetics of these species, it is now clear that the reported
centres [3], these isopoly MBs are very intensely coloured with ‘instability’ of the colour arises from the slow, ongoing dispropor-
molar absorptivities in the order of 105 L mol1 cm1 [118]. Inter- tionation of PMB(2e) to 12-MPA and PMB(4e), which exhibits a
estingly, one study has exploited the apparent catalytic effect of different absorption spectrum. However, once formation of
nanomolar P concentrations on the rate of isopoly MB formation PMB(4e) using organic reductants is complete, the product appears
when Mo(VI) concentrations are much larger and acidities are to be stable for at least several hours with respect to oxidation by
lower than in conventional PMB methods [111]. However, this ef- dissolved O2 [49,60,72,73,124]. Sb(III) methods behave in the same
fect was only apparent with the use of high concentrations of manner; once the reduction finishes after several minutes, the
ascorbic acid; hydrazine sulfate and SnCl2 did not yield the same Sb2PMB(4e) adduct appears stable for many hours with respect to
results. oxidation [51,85]. By contrast, a recent study on the use of UV photo-
Secondly, isopoly MB species are capable of aggregating in so- reduction of 12-MPA found that PMB(4e) immediately began to re-
lution to form colloids, with aggregate sizes of at least tens of oxidise once irradiation ceased, as no reducing reagents other than
nanometres [3,120], as well as adsorbing to surfaces due to their PMB existed in the absence of UV light [125].
very large surface area [120]. This is obviously of concern for SnCl2 methods are well known to yield comparatively short-
spectrophotometric methods due to light scattering, but the pro- lived products. However, this unusually short lifespan cannot be
cess of aggregation is a slow one and is generally only problematic attributed solely to re-oxidation by O2, since electrochemical
when long reaction times are used. Of greater importance is the measurements of Sn-substituted PMB complexes show that these
formation of coatings and deposits on the inside surfaces of flow products are in fact more stable toward oxidation than PMB(4e)
analysis systems where automated MB methods are implemented. [96]. Instead, it has been found that the Sn-substituted products are
However, the principles underlying many modern methods (short hydrolytically unstable and are degraded in a stepwise fashion,
E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82 73

Table 5
Chemical parameters and linear ranges of Sb(III) MB methods.

Reagent concentrations at linear range maximum Ref.

Author (year) [Hþ] (mmol L1) [Mo(VI)] (mmol L1) [AA] (mmol L1) [Sb] (mmol L1) [P] (mmol L1) [Sb]/[P] [Mo(VI)]/[P]

Harwood (1969) 200 2.7 23 329 >97 Excess <28 [83]


Edwards (1965) 57 1.5 1.4 149 71 2.1 21 [84]
Murphy (1962) 200 5.4 4.8 66 32 2.0 167 [85]
Drummond (1995) 127 2.4 96 58 ~26 2.2 93 [29]

presumably by acidified Mo(VI) [96], which may proceed to ulti- The addition of further Sb(III) does not accelerate the reduction
mately form PMB(4e). However, in comparison with heated and process, nor does it appear to interfere with the reduction chem-
Sb(III)-containing ascorbic acid methods, the reductant concen- istry, but the extension of the linear range to arbitrarily high P
tration in SnCl2 methods is normally much lower, which presents a concentrations by adding more Sb(III) is limited by solubility
greater risk of untimely product re-oxidation. Therefore, SnCl2 problems, which have been reported at working concentrations of
methods are often augmented with a sacrificial co-reductant such ca. 104 mol L1 Sb(III) [83,85]. A similar phenomenon is well-
as hydrazine sulfate to extend the oxidative stability of the product known during the preparation of the mixed reagent solution in
to 30e60 min [49,68,126]. methods derived from that of Murphy and Riley [85], in which
turbidity may become evident once both Sb(III) and Mo(VI) have
3.4. Method linearity been added to the solution [84,98], and for which the general
recommendation is to keep mixing the solution until the precipi-
The linear range of a given MB method is the range of P con- tate re-dissolves. This phenomenon is curious given the much
centrations over which 12-MPA formation and reduction occur to greater solubility of potassium antimonyl tartrate in pure water
the same degree, and in which no deviation from the Bou- than in combined reagent solutions. In seems likely that the
guereBeereLambert law is observed. MB method linear ranges, insoluble species being formed is in fact a salt of Sb(III) and
particularly those based on the Murphy and Riley method [85], molybdate [127] which forms at low [Mo(VI)]/[Sb(III)] ratios, rather
typically extend up to around 1 mg L1 P. In practice, modern than a basic salt as suggested by Murphy and Riley [85]. If a salt of
methods typically achieve quantitative, or at least substantial Sb(III) and molybdate is responsible for the Sb(III) solubility issues,
reduction, even in the short timeframes of FIA measurements; it is expected that pH manipulation will be ineffective in dissolving
SnCl2 is well-known to reduce 12-MPA almost instantly, and Sb(III) it [127]; dilution should be the preferred course of action.
methods can achieve full reduction within one minute if the Data on the linear ranges of MB methods other than those uti-
chemical conditions are carefully chosen [29]. lising Sb(III) have seldom been reported; many authors claim that
The actual extent of the linear range is governed by the presence their method demonstrates a linear response up to at least a given P
of sufficient acidified Mo(VI) to stabilise 12-MPA, sufficient reduc- concentration, rather than examining the maximum tolerable P
tant, and in the case of antimony methods, sufficient Sb(III). concentration. As an approximation, non-Sb(III) methods appear to
Furthermore, linear ranges reported for FIA/SIA and batch/DA/SFA be limited by reductant concentration as long as [Mo(VI)]/[P]  21.
versions of the same method are not typically comparable, as they For SnCl2 methods, it appears that an approximately 8-fold
treat the sample volume in a fundamentally different way. Whereas SnCl2:12-MPA excess is required for complete product formation;
sample aliquots are diluted to a fixed volume in batch, DA and SFA the standard SnCl2 method deviates from linearity around this
procedures, they undergo dispersion in FIA and SIA. Since the point, even though Mo(VI) is present in more than an 80-fold molar
practical dispersion coefficient of a flow system describing the excess over orthophosphate [98]. Similarly, El-Shamy and Iskander
extent of sample dilution is unknown unless specifically charac- found that the reduction of 12-MPA continued up to a 5- to 6-fold
terised, it is easier to discuss method linearity in the context of molar excess of SnCl2 over 12-MPA [40]. In methods using heating
batch methods. with organic reductants, 10-fold excesses of hydrazine sulfate and
Antimony-based methods are the simplest case. As discussed 1-amino-2-naphthol-4-sulfonic acid (ANS) have also been reported
above, these form a complex in which the ratio of Sb:P is 2:1, and it as sufficient for complete reduction of 12-MPA [65,73], as the
is clear from a number of investigations [29,83e85] that the upper reduction inherently becomes much more favourable and rapid at
end of the linear calibration range of each method occurs at P higher temperatures.
concentrations where the Sb:P ratio becomes less than 2, with the The ionic strength of the sample matrix is not typically prob-
graph suddenly adopting a shallower gradient (Table 5). At this lematic for the linearity of MB methods, as reagent concentrations
point, the [H4PMo12O40]3 product forms which absorbs at 820 nm, are normally sufficiently high to mask any activity effects. For
contributing only slightly to the absorbance increase at 880 nm. By example, Murphy and Riley observed deviations in the analytical
increasing the Sb(III) concentration used in the Murphy and Riley signal of less than 1% when their method was performed in a matrix
method fivefold, Harwood et al. were able to greatly extend the of approx. 0.44 mol L1 Cl [85] as opposed to distilled water. A
linear range of the former [83], whilst simultaneously halving the more extreme case has been reported by Zhang et al. [128] in which
unnecessarily high Mo(VI) concentration. It has been found that a sample matrix containing 1 mol L1 NaCl after a sequential
even with a [Mo(VI)]/[P] ratio as low as 21, Sb(III) still acts as the extraction procedure did attenuate the linear range, but this
limiting reagent (Table 5) [84], suggesting that a 21-fold [Mo(VI)]/ vulnerability can likely be attributed to the very low working
[P] ratio is sufficient for complete 12-MPA formation. Ascorbic acid concentrations of all reagents used.
concentrations of 5e100 mmol L1 are reported (Table 3), with
higher concentrations accelerating the reduction process some- 3.5. Choice of acid
what [29,81]; it has been suggested that ascorbic acid must be
present in a 20-fold excess over 12-MPA to completely reduce it, The MB reaction requires a strong acid, with reported pH values
with or without Sb(III) [81]. generally below 1 to ensure appropriate Mo(VI) speciation and
74 E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82

inhibition of direct Mo(VI) reduction. However, the choice of acid formation is unclear, it can be assumed that due to the low forma-
can negatively impact the reaction; oxidising acids interfere with tion constant of this complex, combined with the high HCl con-
the reduction process, and anion interactions with Mo(VI) species centration required for it to form, the use of hydrochloric acid in the
can perturb the formation of 12-MPA. MB method should not perturb 12-MPA formation in any practical
Neither nitrate nor perchlorate appear to coordinate to Mo(VI) sense and further investigation of HCl as an alternative to H2SO4
species to any significant degree, even in 0.5 mol L1 Mo(VI) so- should be performed. However, if SnCl2 is preferred as the reductant,
lutions in 2.0 mol L1 HNO3 or HClO4, as determined by 95Mo NMR sulfuric acid should be used as the acid instead of hydrochloric acid
and Raman spectroscopic measurements [129]. However, the oxi- due to the salt error, which is discussed in Section 4.2.3.
dising ability of nitric and perchloric acids should discourage their
use in MB methods; even 0.1 mol L1 HNO3 interferes with SnCl2 4. Interferences
reduction of heteropoly acids [114] and HClO4 has been reported to
form a precipitate with SnCl2 and partially oxidise hydrazine sulfate 4.1. Additive interferences
[74].
Literature reports of interactions of sulfate with heteropoly acid 4.1.1. Arsenate
systems are varied [130,131], including one 1915 report of a Just as PO4 3 forms a reducible heteropoly acid, so too does
molybdosulfate species co-existing with 12-MPA, apparently AsO4 3 . The spectral profiles, molar absorptivities, chemical prop-
precipitated as [(NH4)2SO4.5MoO3] [132]. The use of sulfuric acid erties and formation conditions of arsenomolybdenum blue species
significantly decreases the rate constant for 12-MPA formation are very similar to their orthophosphate counterparts (Table 4). As a
compared with nitric or perchloric acids [30]. It is also clear from result, resolution of the two compounds in the same solution, even
UVevisible spectroscopic comparisons that at equilibrium, 12-MPA using multiple wavelength spectrophotometry, is all but impossible
in H2SO4 is markedly more dissociated into its precursor molyb- with satisfactory accuracy. As(V) interference is generally coun-
dates than in HNO3 [27] or HClO4 [30], particularly at pH < 1.0, with tered either by exploiting subtle kinetic and spectral differences as
the effect also induced by addition of sulfate or bisulfate [27]. the two MB species form [135], or by reducing As(V) to As(III)
Spectrophotometric measurements of dilute, monomeric Mo(VI) [136e138] which cannot form a heteropoly acid due to its lack of
solutions have shown that an equilibrium system between tetrahedral geometry. The reduction of As(V) to As(III) is typically
molybdic acid and hydrogen sulfate does indeed exist, involving a performed using sulfur-containing reductants such as dithionite
1:2 reaction between Mo(VI) and HSO4  (pKa ¼ 1.9) [133] to form a (S2 O4 2 ), thiourea ((NH2)2CS) or thiosulfate (S2 O3 2 ), but even so,
deprotonated product. Only the 1:2 complex was reported, the reduction is a slow process which requires heating if comple-
although ESI-MS data have suggested that a number of other tion within tens of minutes is to be achieved [136,138]. The use of
complexes also exist at various Mo(VI) concentrations [20,106]. thiosulfate also requires a source of SO2 such as acidified meta-
The formation of 12-MPA is therefore inhibited to some extent bisulfite (S2 O5 2 ), and the loss of this toxic gas from solution results
in methods using H2SO4 (or acidified sample matrices containing in precipitation of colloidal sulfur [138] which is particularly trou-
SO4 2 ) since this reaction must involve the release of HSO4  from blesome for MB methods, since they depend on spectrophoto-
molybdosulfate species into a solution already containing a high metric detection.
concentration of HSO4  . This conclusion correlates very well with Multiple authors have observed a ‘synergistic’ effect between
the early observations of Crouch and Malmstadt [27], whereby the As(V) and P in solution, such that solutions containing both appear
extent of 12-MPA formation in H2SO4 was seen to be far less than to form a molybdenum blue species much more quickly than either
that observed with HNO3 at a range of pH values, and Linares et al. of them in isolation [135,137,139,140]. In methods using Sb(III), the
[134] have reported that even using HNO3 instead of H2SO4 results absorption maximum of the new product appears to be redshifted
in a significant enhancement in PMB(4e) formation, despite the by 10 nm compared with the superposition of AsMB and PMB
oxidising power of HNO3. A more detailed investigation of the spectra in isolation [137]. Johnson and Pilson demonstrated that at
impact of H2SO4 on MB method sensitivity is warranted. their measurement wavelength of 865 nm, the final absorbance of a
It is striking, then, that the vast majority of published MB mixed As(V)/P solution was lower, but attained much more rapidly,
methods utilise sulfuric acid. This can potentially be attributed to than the same solution without added P [137]. These authors also
following in the footsteps of Murphy and Riley's seminal contri- showed evidence that the species responsible for this synergistic
bution published in 1962 [85], which itself may have used sulfuric behaviour was a molybdenum blue species containing a 1:1 ratio of
acid as a means of avoiding the formation of iron chloro-complexes As(V):P. Dhar et al. [140] have since provided further evidence for a
in hydrochloric acid, as did a number of its predecessors [93]. It is combined As(V)/P product in their As(V) method; in the presence of
also likely that this acid's popularity is due to consistently low sufficient P, the reduction process was complete in 10 min, but in
levels of PO4 3 contamination, even in analytical reagent grade, natural waters containing < 2 mmol L1 P, the reaction required
thus eliminating the need for more expensive ultra-pure acid. In about 45 min for complete reduction to Sb2AsMB to occur. Aside
some samples such as peroxydisulfate digests, SO4 2 and HSO4  from the substantial kinetic effect of this phenomenon, however,
are unavoidable in any case. Furthermore, the ‘salt error’ encoun- the effect on the final absorbance of the solution typically gives rise
tered in SnCl2 methods appears to have discouraged the use of to an error of only ~5%. Furthermore, these authors made the
hydrochloric acid as an alternative to sulfuric acid, even in cases intriguing observation that increasing the concentration of reduc-
where the Cl ion should not cause any interference. tant (ascorbic acid) served to drastically decrease the impact of this
Like sulfate, chloride is also known to form complexes with combined product on the final absorbance [140]. Lo pez Carreto
molybdate cations, and the complexes MoO2Cl2(H2O)2 and MoO- et al. observed similar synergistic phenomena with the standard
Cl3(OH) (H2O) have been identified [25,129]. However, these com- PMB(4e) product [135]. If a discrete species containing both As
plexes are much weaker than similar sulfate complexes, with and P is indeed formed under the conditions used by any of the
formation constants of 0.29 and 0.036 respectively. Additionally, the aforementioned authors, its composition is unclear, but it may be a
formation of chloro-complexes with molybdate cations is only seen P-substituted derivative of an arsenomolybdate complex, given the
at HCl concentrations higher than 0.4 mol L1, and approximately diverse range of As:Mo stoichiometries encountered in these
0.8 mol L1 HCl is required for 10% of the molybdate species to form compounds [35]. Whether or not the formation of combined As(V)/
the dichloro-complex. Whilst the effect of this species on 12-MPA P species occurs in SnCl2 reduction methods is unknown.
E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82 75

4.1.2. Silicate that the tartaric acid acts to suppress silicate interference, even in
SiO4 4 has long been considered one of the main interferents in micromolar quantitites. Zhang et al. have reported that the
PMB methods as it also forms heteropoly acids (a- and b-12-MSA) addition of this salt can reduce silicate interference by almost 50%
[114] reducible to molybdenum blues. Silicate interference is well- when the method is heated to 70  C [109], although whether this
known to become problematic when the reaction acidity is too low result can be attributed to the presence of tartaric acid or the
and when heating is employed [29,109]. This is due to two separate formation of Sb-containing heteropoly acids is uncertain.
phenomena; 12-MSA is only stable at lower Z values (higher pH)
than 12-MPA [73], and the speciation of silicic acid itself is 4.1.3. Organic and inorganic P hydrolysis
dependent on both temperature and acidity. At low pH, orthosilicic Concern is frequently voiced about the undesirable tendency
acid (H4SiO4) exists in equilibrium with polysilicic acids [141,142] of MB methods to hydrolyse other fractions of total dissolved
which form 12-MSA much more slowly than orthosilicic acid phosphorus (TDP), thereby providing an overestimate of ortho-
alone [142], presumably because only orthosilicate possesses the phosphate. This is the basis for labelling the phosphorus fraction
appropriate molecular geometry to form 12-MSA. Of note is a very determined by the MB method as 'dissolved reactive phosphorus'
recent ESI-MS study [20] suggesting that the use of H2SO4 actually (DRP) or, more correctly, ‘molybdate reactive phosphorus’ (MRP)
decomposes polysilicic acids into a monosilicate complex with [2], which is assumed to also include some proportion of inor-
HSO 4 which may still react to form 12-MSA, implying that silicate ganic polyphosphates, labile organic P compounds and colloidal
interference is therefore increased when H2SO4 is used. Unfortu- P species [2]. Hydrolytic degradation of phosphates is quite
nately, no more data on this phenomenon are available at this time. distinct from oxidative processes, which are only introduced in
Higher reaction temperatures favour the decomposition of methods for TDP determination. Whilst (thermal) acid hydrolysis
polysilicic acids into orthosilicic acid [142], a phenomenon which is effective for conversion of polyphosphates and labile organic P
probably gave rise to silicate's reputation as a major interferent in compounds, oxidative processes are much more effective for the
PMB methods due to the frequent use of heating employed in older total dissolved organic phosphorus (DOP) pool [50,148e153]; as
literature methods (Table 3). Higher reaction temperatures also such, additive MB interference is concerned mainly with the
broaden the Z range in which both isomers of 12-MSA will form, former.
exacerbating the problem even further [114]. However, it has been Hydrolysis and desorption of other P-containing fractions in TDP
reported that heated acid digestion procedures are capable of re- is widely considered to be acid-induced. However, a number of
polymerising silicic acid into unreactive species through dehydra- studies have shown that both acid and Mo(VI) act to hydrolyse
tion [143]. these species [71,154e156], and that the influence of Mo(VI) in such
Due to the phenomena described above, silicate interference can cases can be pronounced [154]. It has been tentatively suggested
be effectively controlled by using a sufficiently high acidity that Mo(VI) acts by binding to phosphate groups and then
[114,141] in addition to avoiding heating, which is rarely used in removing them from the parent compound [154,156], which is
more modern PMB methods (Table 3). The rate of 12-MSA forma- supported by the extremely high affinity of Mo(VI) for orthophos-
tion decreases sharply when Z > 2 [90], which can be approximated phate discussed earlier [32].
as > 0.2 mol L1 Hþ for low millimolar Mo(VI) concentrations and is
within the optimal range for many PMB methods (Table 3). For 4.2. Subtractive interferences
example, in a previously reported flow injection method where
working concentrations of 5 mmol L1 Mo(VI) and 4.2.1. Organic acids
0.2 mol L1 H2SO4 were used, 50 mg L1 Si did not interfere at all The presence of organic acids is known to inhibit the MB re-
and 100 mg L1 Si generated only a þ2.6% error in the timeframe of action, particularly in the context of masking one oxoanion in
the method [125]. Since acidity controls silicate interference by order to more accurately determine another. The nature of organic
slowing the kinetics of 12-MSA formation, it therefore follows that acid interference in PMB methods is the formation of coordination
the more rapidly a measurement is made, the more silicate can be complexes with Mo(VI) [144], acting to both sequester Mo(VI) and
tolerated in a sample without detectable interference [109]. destroy 12-MPA [145]. However, this can only occur if the species
An alternative approach to inhibiting 12-MSA formation is the in question is able to coordinate in a bidentate manner to form a
use of organic acids. These species function as ligands, seques- stable 5- or 6-membered coordination ring with Mo(VI) [157]. A
tering Mo(VI) and thus slowing the formation of 12-MSA, a number of Mo(VI) complexes with common organic anions have
phenomenon further discussed in Section 4.2.1. This approach been characterised, such as tartrate [158], oxalate [159e161],
exploits the observation that 12-MSA's formation kinetics can be malate [162], and citrate [163,164], and these complexes persist at
greatly slowed by the presence of organic acids in comparison to the acidities typically used in MB methods. Formic and maleic
12-MPA which forms much more rapidly. However, this approach acids do not appear to interfere with orthophosphate determi-
can be a double-edged sword since 12-MPA is decomposed by nation [165]; these species lack the necessary atom(s) on an a-
such species at higher concentrations, whereas 12-MSA is unaf- carbon to create a 5- or 6-membered coordination ring with
fected by organic acids if they are added after its formation is Mo(VI), and thus coordinate too weakly to be problematic. In fact,
complete, a phenomenon often exploited in silicate determina- organic acids are sometimes used in MB methods to prevent the
tion [144]. For example, 12-MSA formation can be effectively reduction of excess Mo(VI), particularly in silicate methods where
suppressed with dilute oxalic acid whilst leaving 12-MPA mostly the requisite acidity is lower than for orthophosphate [166,167].
intact, whereas fully formed 12-MPA can be easily destroyed with By the same token, interference from organic ligands in PMB
more concentrated oxalic acid, which is ineffective in decom- methods can be eliminated by complexing these species with
posing fully formed 12-MSA [145]. Tartaric acid (C4H6O6) is also excess Mo(VI) [147].
often used for inhibiting silicate interference in orthophosphate
determinations [145,146], yet it is reported to interfere with this 4.2.2. Fluoride
determination quite substantially in its own right [145,147], more F is a strong negative interferent in the MB reaction
so than oxalic acid [145]. Interestingly, since methods using [68,147,168,169]. In a similar manner to that of organic acids, its
Sb(III) almost invariably utilise potassium antimonyl tartrate mode of interference is binding to Mo(VI) to form discrete ions
(K2Sb2(C4H2O6)2.3H2O) as the source of Sb(III), it may be expected which cannot condense to form larger structures such as 12-MPA
76 E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82

[170,171]. This has the effect of slowing the MB reaction without the method can be calibrated in that matrix (e.g. seawater) or a
necessarily changing the extent of the reaction, giving the solution with matching salinity. The third option is to heat the re-
appearance of a lower analytical signal if insufficient time is action to near-boiling for several minutes in an analogous way to
allowed for the inhibited reaction to reach completion [172]. H3BO3 heated ascorbic acid methods; this results in the exclusive forma-
is very effective at sequestering F as BF4  without negatively tion of PMB(4e) by accelerating both its direct formation and the
impacting the MB reaction [170,173]. An interesting point is that hydrolysis of SnPMB(4e) [62], thus eliminating the role of Sn(IV) in
silicate also complexes to F in solution, and can mitigate the the reaction entirely [97]. However, the use of heating in this
inhibitory effect of F on the MB reaction in its own right [101]. Of approach makes it highly vulnerable to silicate interference.
course, this phenomenon interferes with the determination of sil-
icate itself in waters containing sufficient F. 4.3. Multifunctional interferents
Interestingly, F can also alter the redox potentials of metallic
reductants. The use of Fe(II) as a reductant combined with NaF has 4.3.1. Sulfide
been reported, wherein the complexation of F to Fe(II) enhanced Sulfide, present as H2S under acidic conditions, is capable of
the latter's reducing power [174]. NaF has also been used to both additive and subtractive interference in the MB reaction [181].
enhance the reducing power of SnCl2 by converting its oxidation H2S can act as a reductant for 12-MPA in its own right [182,183] and
product to the more stable SnF6 2 [175]. However, due to the for- causes distinct additive interference at low P concentrations as a
mation of HF at the acidities used in the MB reaction, this approach result [181], particularly in terms of direct Mo(VI) reduction. At
cannot be recommended on practical grounds. higher P concentrations, H2S causes subtractive interference
instead, apparently by complexing with Mo(VI) [181]. In Sb(III)
4.2.3. Chloride (salt error) methods, H2S will also sequester Sb(III) to form insoluble antimony
A major source of subtractive interference in MB methods is the sulfides; not only is the precipitate problematic for spectropho-
'salt error' [50,85,97,105,176], which manifests as a decrease in tometry, but the loss of dissolved Sb(III) also forces the formation of
analytical signal when SnCl2 is used as the reductant in assays of the PMB(4e) species, thus altering the spectral properties of the
marine and estuarine waters or other matrices containing elevated product [184]. H2S can be conveniently oxidised by KMnO4 prior to
chloride concentrations. Whilst a small (approx. 1e4%) decrease in the MB reaction [183], provided that excess reductant is subse-
analytical signal is generally expected at high ionic strengths due to quently used, or it can be removed by acidification and subsequent
the decrease in activity coefficients [50,176], the decrease in signal degassing [181]. Concerning the former approach, a working
is usually reported to be between 15 and 20% in marine samples KMnO4 concentration of 5.3 mmol L1 can be reduced by a working
[50,64,85,176] and even higher in some cases [68,97], implying a ascorbic acid concentration of 26.9 mmol L1 with no impact on
dependence on method conditions as well. The degree of analytical method sensitivity [183].
signal suppression varies with the Cl concentration in the matrix
[68,97] as well as the orthophosphate concentration [97], and 4.3.2. Iron
method linear range is attenuated with increasing Cl concentra- Fe(II), much like H2S, can reduce 12-MPA. In fact, Fe(II) has been
tion. As such, the salt error is troublesome not only because of the used as the main reductant in several MB methods [66,185] as well
loss of method sensitivity, but because this loss varies in magnitude as the analyte by virtue of its reducing ability [186]. However, this is
between sample matrices of different salinities and P only troublesome in terms of method interference if the reductant
concentrations. concentration is too low or the acid e molybdate balance favours
The nature of the salt error is the decreased formation of isopolymolybdenum blue formation. Fe(III) is more problematic as
SnPMB(4e) due to the disruptive effect of Cl on the chemistry of it appears to consume the reductant, reducing sensitivity and/or
Sn(IV). Chloro-complexes of Sn(II) and Sn(IV) readily form in the product stability [64,187e190]. It has also been reported that Fe(III)
presence of Cl [105,177e179], and their stability dramatically might precipitate orthophosphate from solution [191]. Fe(III)
impedes the ability of Sn(IV) to substitute into PMB(4e) to form interference is effectively controlled with a sufficient excess of
Sn2PMB(4e) [97,105]. As a result, the highly favourable combined reductant or the presence of an organic acid to complex with Fe(III)
reduction and substitution step responsible for producing [188,189,192].
Sn2PMB(4e) is impeded due to the depletion of uncomplexed
Sn(IV). Thus, higher concentrations of Cl limit the extent to which 4.3.3. Surfactants
SnPMB(4e) can form, and force the system to reduce 12-MPA via It has been reported that different types of surfactants interact
the slower and much less favourable direct reduction process [97]. with the MB reaction when using either ascorbic acid/Sb(III) or
This is clearly shown in UVevisible spectra of products suffering SnCl2 reduction [193]. In this study, cationic surfactants such as
from the salt error, in which the 710 nm SnPMB(4e) peak is tetraalkylammonium chlorides were shown to cause severe nega-
diminished and a distinct shoulder at 820 nm emerges due to the tive interference even at low mg L1 concentrations; 5 mg L1
increased presence of PMB(4e). Even though the two reduction surfactant caused 15% sensitivity loss for both reduction methods,
products possess absorption maxima 100 nm apart, dual- and higher surfactant concentrations caused turbidity. A neutral
wavelength spectrophotometry cannot be used to counter the salt detergent species, nonylphenol ethoxylate, also induced turbidity
error since the total concentration of reduced products also de- at similar concentrations; however, its effect on the absorbance
creases with higher Cl concentrations [97]. resulting from ascorbic acid/Sb(III) reduction was variable, whilst
Three strategies for countering the salt error have been pro- the SnCl2 procedure seemed largely unaffected despite the
posed. Firstly, the reaction can be performed using reagents which turbidity. By contrast, commonly used anionic surfactants such as
already contain a sufficiently concentrated Cl ‘buffer’, such that linear alkyl sulfonates did not interfere at any of the tested con-
the sample matrix does not perturb the Cl concentration at the centrations for either reduction method (up to 2.5 g L1 for ascorbic
time of reduction [180]. This approach does not eliminate the acid/Sb(III) reduction), and branched sulfonates such as sodium
sensitivity loss caused by the salt error but it does buffer against dioctylsulfosuccinate only interfered above 25 mg L1. Neither of
variations in the magnitude of the salt error between samples, these anionic species caused any turbidity. Industrial detergent
allowing a single calibration curve to be used for analysis. Secondly, formulations were found to cause interference through other
if a large number of samples in similar matrices are to be analysed, components of the mixture in addition to the surfactants
E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82 77

themselves, though again only above 25 mg L1. interior of the flow manifold. This effect often manifests as a long
Interference from neutral surfactants can be completely elimi- tail on a signal peak as the deposited product is washed out of the
nated in either reduction method by adding a linear alkyl sulfonate photometer flow cell, a return to a non-zero absorbance baseline
in a 10-fold excess over the interfering species [193]. However, after the passage of an analyte peak, or an irregular, positive y-
interference from cationic surfactants was only able to be elimi- intercept in batch or DA analysis formats, an occurrence described
nated in the SnCl2 procedure, requiring a 5-fold excess of linear as sample ‘carryover’ [109,204]. Coating reduces the sample
alkyl sulfonate over the cationic surfactant. In light of the above, it is throughput of a method due to the extra time required for washing
probable that the severe interference and precipitation caused by the system, and compromises the precision of subsequent analyses
cationic surfactants is due to the formation of ion-pairs with if sample carryover is still present in the system.
molybdate or PMB anions, whereas the interference from the other The coating phenomenon is dependent on the pH of the reaction
types of surfactants may be due to aggregation behaviour. The use mixture and the concentration of PMB product formed. The general
of linear alkyl sulfonates evidently sequesters cationic surfactants observation that coating causes a smooth, elongated tail on FIA
and/or disrupts their interactions with the reacting species. In fact, peaks suggests that the coating species exhibits a similar or iden-
the use of anionic surfactants in MB methods has previously been tical spectral profile to the PMB product being measured. Combined
reported for different reasons; sodium dodecyl sulfate has with the observation that higher concentrations of PMB product
frequently been used in automated methods to minimise the increase sample carryover [109], it is apparent that the coating
deposition of MB species on the surfaces of cuvettes or flow cells compound, which appears blue [109,126], is the main reduction
[99], an application discussed at length in Section 5. product and not the isopoly MB species formed in the reagent
blank. The effect of acidity on sample carryover is interesting; for
5. MB chemistry in flow methods PMB(4e), the amount of sample carryover appears to reach a
minimum at pH 0.5 and varies little at higher pH, but is greatly
The MB reaction is finding increasing use in flow analysis increased below pH 0.5 [109,195]. By contrast, Sb2PMB(4e) ex-
methods for P determination, taking advantage of the superior hibits increased coating at pH values both above and below 0.5
sampling rate and reproducibility inherent to these techniques [195], with precipitation of Sb(III) salts above pH 1.50 [109]. The
when compared with batch methodologies [2,194,195]. However, coating phenomenon has previously been attributed to the
beyond the inherent benefits of automation, flow analysis is often colloidal behaviour of PMB species [109,195] caused by their limited
used as a platform for coupling P determination with more so- water solubility. This explanation is reasonable given the clear
phisticated operations which would be tedious or impossible to decrease in coating incidence at higher temperatures, and suggests
perform reproducibly under batch conditions. These include the that the solubilities of both PMB(4e) and Sb2PMB(4e) decrease
determination of different P fractions using UV photo-oxidation or dramatically below pH 0.5, whilst Sb2PMB(4e) also experiences
acid hydrolysis [148e150,196], the use of reaction kinetics for P solubility problems at higher pH values. The actual reductant used,
determination [197], independent determination of both P and As so long as the reduced product is the same, has no inherent effect
fractions [134], preconcentration [99,198] and UV photo-reduction on coating; assertions to the contrary are very likely due to differing
of 12-MPA using ethanol, allowing the use of a single long-lived extents of reduction between methods [195].
reagent solution [125]. The lower solubility of the Sb2PMB(4e) complex than that of
Flow analysis manifolds with spectrophotometric detection can PMB(4e) has been investigated by Zhang et al. in two studies
incorporate liquid-core waveguide flow cells, which increase the [109,195]. These authors showed that between pH 0.5e1.5,
optical path length (up to 1000 cm) by the use of total internal Sb2PMB(4e) gave rise to a carryover percentage of between 1 and
reflection without significantly attenuating the light beam 2% at room temperature, roughly double that of PMB(4e) [109].
[199,200]. This technology potentially allows for significant im- Curiously, these results differ greatly from those obtained in a
provements in MB method sensitivity, although several practical subsequent study by the same authors in which carryover co-
issues such as back-pressure, bubble formation, the Schlieren effect efficients were significantly larger [195]. However, this later study
and signal noise must be carefully managed [146,199,201,202]. The also used a commercial anionic surfactant formulation with un-
Schlieren effect is of particular concern in saline samples, which known additives; the potential precipitation problems associated
often demand the greatest sensitivity of any P determination with such formulations have been previously discussed in Section
method. To this end, reverse FIA has been used for the analysis of 4.3.3.
samples with variable salinities to effectively counter the Schlieren Fortunately, the coating of PMB products is straightforward to
effect [203]. control. It is recommended that either a low working concentra-
In terms of actual reaction chemistry, the Murphy and Riley tion of glycerol (3.5e5.0% v/v) [126,134] or sodium dodecyl sulfate
method [85] using ascorbic acid/Sb(III) is by far the most popular (0.05% m/v) [99] should be added to the reaction mixture to
of any MB method in flow analysis owing to its relatively fast suppress the deposition of reduction product. The coating
reaction kinetics at room temperature, insensitivity to Cl inter- behaviour of SnPMB(4e) has not been studied, but the precau-
ference and low susceptibility to silicate interference, since heat- tionary use of sodium dodecyl sulfate (0.05% m/v) has been
ing is not required [2]. SnCl2 methods also have considerable effective in suppressing coating caused by this product as well
utility in a flow analysis setting; their very rapid reduction kinetics [99]. The critical micelle concentration of sodium dodecyl sulfate
and more intensely absorbing product [97] allow for greater varies considerably depending on solution composition [205], and
sensitivity and sample throughput than ascorbic acid/Sb(III) a concentration of 0.05e0.20 % (m/v) of this surfactant in the final
methods, since their limited stability after several minutes is reaction mixture is recommended to ensure that the critical
typically of little concern. However, the salt error has led to the micelle concentration is exceeded. The effect of high acidity on
diminished popularity of such methods, particularly as many flow PMB solubility is generally not problematic since a pH < 0.5 is
analysis applications are aimed at ultra-trace P determination in expected to be highly detrimental to MB method sensitivity in any
marine or estuarine waters. case, as discussed earlier in this review. Coating can be decreased
An important phenomenon prevalent in flow-based, batch and by using a higher reaction temperature, as would be expected,
discrete analyser formats alike is that of ‘coating’, or the deposition although any such benefits for systems producing PMB(4e) are
of reaction products on the walls of the reaction vessel or the typically small [109].
78 E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82

6. Conclusions and recommendations sensitive to small changes in acidity, whilst using much larger
Mo(VI) concentrations than are necessary increases the acidity
The MB reaction consists of multiple interacting equilibria based needed to suppress blank formation, which decreases sensitivity.
on the complex speciation of aqueous Mo(VI), the formation of The acidity should then be varied such that the blank is minimised
phosphomolybdic heteropoly acids and the reduction of 12-MPA by but sensitivity is not reduced. As an example, it has been previously
various organic or metallic species. Several long-standing as- shown that 3e5 mmol L1 Mo(VI) is more than sufficient for
sumptions about this reaction have been shown to be incorrect and complete 12-MPA formation from 1 mg L1 (32 mmol L1) P, and
counterproductive for method optimisation. In particular, it is that an acid concentration of around 0.20 mol L1 Hþ is an effective
demonstrated that the concept of the [Hþ]/[Mo(VI)] ratio, first match for this Mo(VI) concentration range. Whether Na2MoO4 or
introduced by Strickland [115] and widely adopted since, is a (NH4)6Mo7O24.4H2O is used is of no consequence for the reaction
parameter which fails to define any chemical property of the MB due to the equilibration of Mo(VI) species in acidic solutions within
system and is an entirely misleading framework with which to tens of minutes. If Sb(III) is used, the Sb concentration must be at
approach MB method optimisation. least twice the highest expected P concentration in order to ensure
Several possible ‘molybdenum blue’ species are identified as that the method's linear range is sufficient; excess Sb(III) does not
end products of the reaction depending on the conditions used. cause interference but high working concentrations may result in
[H4PMo12O40]3 is the reduction product in methods which use precipitation. Potassium antimonyl tartrate is a suitable source of
organic reductants and/or heating, which may coexist with Sb(III), and at typical working concentrations of ca. 105 mol L1,
[H3PMo12O40]4 in methods using very low acidities. In contrast, the interference of tartaric acid should be insignificant. The optimal
the use of Sn(II) or Sb(III) in the reduction step yields MB species reductant concentration will vary depending on the type of
incorporating these metals. Each of these MB species discussed reducing system used (organic reductant with heating, organic
above can be identified by its own distinctive Visible-NIR spectra. A reductant in the presence of Sb(III), or Sn(II)) and can be simply
mixture of giant isopolymolybdenum blues constitutes the blank optimised for a given pair of Hþ and Mo(VI) concentrations by
signal of any MB method, which develops more readily with higher determining the concentration of reductant at which sensitivity is
Mo(VI) concentrations and lower acidities. maximal but blank formation does not occur. Typical reductant
The following practices are recommended in the use of the MB concentrations are approximately 1 mmol L1 SnCl2 or
reaction for P determination: 5e20 mmol L1 ascorbic acid (Table 3). Silicate interference is easily
controlled with sufficient acidity (pH < 1), short reaction times and
6.1. Recommended reductants the avoidance of heating. All MB methods should use a means of
minimising product coating; working concentrations of glycerol
In general, ascorbic acid and Sb(III) should be used since the (3.5e5.0% (m/v)) or sodium dodecyl sulfate (0.05% (m/v) or higher)
Sb2PMB(4e) reduction product forms within minutes, is stable for have proven effective. This approach is recommended for flow and
hours and is insensitive to chloride interference. This is particularly batch methods alike.
the case in batch, DA and SFA methods, where the temporal sta- The authors of the present work recommend against the use of
bility of the product is of greater importance. However, when the [Hþ]/[Mo(VI)] ratio in method optimisation as it is a chemically
maximum sensitivity is desirable and sample matrices contain unjustifiable and entirely empirical variable which is not a good
negligible chloride concentrations, SnCl2 in combination with a indicator of MB reaction chemistry. Amongst the literature
sacrificial co-reductant (typically N2H6SO4) should be used in methods using Sb(III) and ascorbic acid, the apparently optimal
preference to ascorbic acid and Sb(III) due to the faster reduction [Hþ]/[Mo(VI)] ratio varies between 37 and 74 (Table 3), even after
kinetics and significantly (~30%) higher molar absorptivity of correcting for the common yet erroneous assumption that H2SO4
SnPMB(4e). fully dissociates twice at pH < 1. Several further points must be
considered:
6.2. Recommended acids
a) The [Hþ]/[Mo(VI)] ratio does not inherently prescribe actual
Since both HNO3 and HClO4 are oxidising acids and interfere to a reagent concentrations, nor can it. Therefore, MB reaction
considerable extent with the reduction process, HCl and H2SO4 are chemistry may differ substantially between two methods using
recommended as the strong acids of choice for MB procedures. different reagent concentrations but the same [Hþ]/[Mo(VI)]
However, there is some evidence to suggest that H2SO4 inhibits the ratio [82]. Reagent concentrations can reportedly be scaled up or
reaction, and a comparative study of H2SO4 and HCl is warranted down for methods using Sb(III) at a fixed [Hþ]/[Mo(VI)] ratio of
for MB procedures. If SnCl2 is used as the reductant, H2SO4 should 35, but only between Mo(VI) concentrations of
be used due to salt error interference from the Cl ion unless it is 0.84e8.40 mmol L1 [81]. However, this lower limit appears to
desirable to use HCl as a means of buffering the Cl concentration. If be due to the P concentration used in these experiments, whilst
converting acid concentrations between H2SO4 and HCl, or when at the upper limit where [Hþ] ¼ 0.588 mol L1, Z is sufficiently
consulting older literature, care must be taken to accurately high for 12-MPA to begin decomposing. The linear range will
calculate acidity, since the older normality scale of acidity is an also be limited by the Mo(VI) concentration used.
inaccurate measure of [Hþ] when a diprotic acid such as H2SO4 is b) If the reaction is heated, silicate interference increases signifi-
used. cantly at lower acidities, even if the [Hþ]/[Mo(VI)] ratio remains
fixed.
6.3. Recommended optimisation procedure c) Reaction time varies significantly, and in a complex manner,
with both [Hþ] and [Mo(VI)] [82]. At a fixed [Hþ]/[Mo(VI)] ratio,
Any MB method for P determination should be optimised by the reaction time invariably increases as the acid concentration
first selecting a Mo(VI) concentration which will be effective in the increases.
desired P concentration range (a 21-fold [Mo(VI)]/[P] excess at the
highest predicted P concentration should be the minimum). It is considerably more complicated to optimise a method based
Caution is required however, since very low (sub-millimolar) on the trial-and-error [Hþ]/[Mo(VI)] framework as opposed to the
Mo(VI) concentrations will make the reaction impractically method described above of selecting an appropriate Mo(VI)
E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82 79

concentration, optimising the acidity on this basis, and then opti- [24] J. Burclova, J. Pr
asilov
a, P. Benes, The state and adsorption behaviour of traces
of molybdenum(VI) in aqueous solutions, J. Inorg. Nucl. Chem. 35 (1973)
mising the reductant concentration. As such, significantly greater
909e919.
clarity in the literature could be achieved if MB methods were to [25] S. Himeno, M. Hasegawa, Spectrophotometric studies on the monomer-
report their working concentrations of Hþ, Mo(VI) and Sb(III) where monomer equilibration of Mo(VI) in hydrochloric acid solutions, Inorg.
appropriate, such as in Table 3, thus enabling at-a-glance compar- Chim. Acta 73 (1983) 255e259.
[26] F. Taube, I. Andersson, L. Pettersson, Molybdate speciation in systems related
ison of the meaningful parameters which underlie Mo(VI) specia- to the bleaching of kraft pulp, in: M.T. Pope, A. Müller (Eds.), Poly-
tion, blank formation, silicate interference, linear range and oxometalate Chemistry from Topology via Self-assembly to Applications,
reaction rate. Springer, Netherlands, 2001, pp. 161e173.
[27] S.R. Crouch, H.V. Malmstadt, Mechanistic investigation of molybdenum blue
method for determination of phosphate, Anal. Chem. 39 (1967) 1084e1089.
Acknowledgements [28] P. Cannon, Some electrometric measurements of heteropoly ion formation in
aqueous systems, J. Inorg. Nucl. Chem. 13 (1960) 261e268.
[29] L. Drummond, W. Maher, Determination of phosphorus in aqueous solution
E. A. Nagul is grateful to the University of Melbourne for the via formation of the phosphoantimonylmolybdenum blue complex. Re-
award of a postgraduate scholarship. P.J. Worsfold is grateful to the examination of optimum conditions for the analysis of phosphate, Anal.
University of Melbourne for the award of a Wilsmore Fellowship. Chim. Acta 302 (1995) 69e74.
[30] P.M. Beckwith, A. Scheeline, S.R. Crouch, Kinetics of the formation of 12-
molybdophosphate in perchloric, sulfuric, and nitric acid solutions, Anal.
References Chem. 47 (1975) 1930e1936.
[31] A.C. Javier, S.R. Crouch, H.V. Malmstadt, Investigations of formation of 12-
[1] P.J. Worsfold, L.J. Gimbert, U. Mankasingh, O.N. Omaka, G. Hanrahan, molybdophosphoric acid utilizing rapid reaction-rate measurements, Anal.
P.C.F.C. Gardolinski, P.M. Haygarth, B.L. Turner, M.J. Keith-Roach, Chem. 40 (1968) 1922e1925.
I.D. McKelvie, Sampling, sample treatment and quality assurance issues for [32] K. Murata, T. Kiba, Studies on the formation and the extraction of molyb-
the determination of phosphorus species in natural waters and soils, Talanta dophosphoric acid, J. Inorg. Nucl. Chem. 32 (1970) 1667e1678.
66 (2005) 273e293. [33] J.A. van Veen, O. Sudmeijer, C.A. Emeis, H. de Wit, On the identification of
[2] I.D. McKelvie, D.M.W. Peat, P.J. Worsfold, Techniques for the quantification molybdophosphate complexes in aqueous solution, J. Chem. Soc. Dalton
and speciation of phosphorus in natural waters, Anal. Proc. Incl. Anal. Trans. (1986) 1825e1831.
Commun. 32 (1995) 437e445. [34] R.I. Buckley, R.J.H. Clark, Structural and electronic properties of some poly-
[3] A. Müller, C. Serain, Soluble molybdenum blues“des Pudels Kern”, Acc. Chem. molybdates reducible to molybdenum blues, Coord. Chem. Rev. 65 (1985)
Res. 33 (2000) 2e10. 167e218.
[4] J.F. Keggin, Structure and formula of 12-phosphotungstic acid, Proc. R. Soc. [35] S. Himeno, M. Hashimoto, T. Ueda, Formation and conversion of molybdo-
Lond. Ser. A 144 (1934) 75e100. phosphate and -arsenate complexes in aqueous solution, Inorg. Chim. Acta
[5] C.G. Bochet, T. Draper, B. Bocquet, M.T. Pope, A.F. Williams, 182Tungsten 284 (1999) 237e245.
Mossbauer spectroscopy of heteropolytungstates, J. Chem. Soc. Dalton Trans. [36] C.C. Kircher, S.R. Crouch, Determination of formation constants of molyb-
(2009) 5127e5131. dophosphates in strong acid solutions, Anal. Chem. 54 (1982) 879e884.
[6] J.J. Cruywagen, A.G. Draaijer, J.B.B. Heyns, E.A. Rohwer, Molybdenum(VI) [37] K.M. Reddy, N. Lingaiah, P.S.S. Prasad, I. Suryanarayana, Acidity constants of
equilibria in different ionic media. Formation constants and thermodynamic supported salts of heteropoly acids using a methodology related to the
quantities, Inorg. Chim. Acta 331 (2002) 322e329. potentiometric mass titration technique, J. Solut. Chem. 35 (2006) 407e423.
[7] K. Murata, S. Ikeda, Studies on polynuclear molybdates in the aqueous so- [38] K. Maeda, S. Himeno, T. Osakai, A. Saito, T. Hori, A voltammetric study of
lution by laser Raman spectroscopy, Spectrochim. Acta Part A 39 (1983) Keggin-type heteropolymolybdate anions, J. Electroanal. Chem. 364 (1994)
787e794. 149e154.
[8] K. Murata, S. Ikeda, Studies on yellow and colourless molybdophosphate [39] J. Gonzalez, A. Molina, M. Lopez-Tenes, F. Karimian, Reversible surface two-
complexes in the aqueous solution by laser Raman spectroscopy, Polyhedron electron transfer reactions in square wave voltcoulommetry: application to
2 (1983) 1005e1008. the study of the reduction of polyoxometalate [PMo12O40]3- immobilized at a

[9] L. Pettersson, I. Andersson, L. Ohman, Multicomponent polyanions. 39. boron doped diamond electrode, Anal. Chem. 85 (2013) 8764e8772.
Speciation in the aqueous Hþ-MoO2 4 -HPO4
2
system as deduced from a [40] H.K. El-Shamy, M.F. Iskander, Studies on some heteropoly bluesdI the
combined Emf-31P NMR Study, Inorg. Chem. 25 (1986) 4726e4733. reduction of 12-molybdophosphoric acid with stannous chloride, J. Inorg.
[10] K. Murata, S. Ikeda, A mechanistic investigation on the formation of Nucl. Chem. 35 (1973) 1227e1237.
molybdophosphate complexes in aqueous solution by the use of laser Raman [41] R.I. Maksimovskaya, Molybdophosphate heteropoly blues: electron-transfer
spectroscopy, Polyhedron 6 (1987) 1681e1685. reactions in aqueous solutions as studied by NMR, Polyhedron 65 (2013)
[11] J.J. Cruywagen, J.B.B. Heyns, Molybdenum(VI) equilibria at high perchloric 54e59.
acid concentration, Polyhedron 19 (2000) 907e911. [42] N. Tanaka, K. Unoura, E. Itabashi, Voltammetric and spectroelectrochemical
[12] J.J. Cruywagen, A.G. Draaijer, Solvent extraction investigation of molybde- studies of dodecamolybdophosphoric acid in aqueous and water-dioxane
num (VI) equilibria, Polyhedron 11 (1992) 141e146. solutions at a gold-minigrid optically transparent thin-layer electrode,
[13] J.J. Cruywagen, J.B.B. Heyns, Spectrophotometric determination of the ther- Inorg. Chem. 21 (1982) 1662e1666.
modynamic parameters for the first two protonation reactions of molybdate: [43] M. Sadakane, E. Steckhan, Electrochemical properties of polyoxometalates as
an advanced undergraduate laboratory experiment, J. Chem. Educ. 66 (1989) electrocatalysts, Chem. Rev. 98 (1998) 219e238.
861e863. [44] A.V. Kolliopoulos, D.K. Kampouris, C.E. Banks, Rapid and portable electro-
[14] K.H. Tytko, G. Baethe, J.J. Cruywagen, Equilibrium studies of aqueous poly- chemical quantification of phosphorus, Anal. Chem. 87 (2015) 4269e4274.

molybdate solutions in 1 M NaCl Medium at 25 C, Inorg. Chem. 24 (1985) [45] A. Hiskia, E. Papaconstantinou, Photocatalytic oxidation of organic com-
3132e3136. pounds by polyoxometalates of molybdenum and tungsten. Catalytic
[15] J.J. Cruywagen, Potentiometric investigation of molybdenum(VI) equilibria at regeneration by dioxygen, Inorg. Chem. 31 (1991) 163e167.

25 C in 1 M NaCl medium, Inorg. Chem. 19 (1980) 552e554. [46] M. Fournier, C. Rocchiccioli-Deltcheff, L.P. Kazansky, Infrared spectroscopic
[16] J.J. Cruywagen, J.B.B. Heyns, E.F.C.H. Rohwer, Dimeric cations of molybde- evidence of bipolaron delocalization in reduced heterododecamolybdates,
num(VI), J. Inorg. Nucl. Chem. 40 (1978) 53e59. Chem. Phys. Lett. 223 (1994) 297e300.
[17] J.J. Cruywagen, J.B.B. Heyns, E.F.C.H. Rohwer, Spectrophotometric investiga- [47] J.J. Borr
as-Almenar, J.M. Clemente, E. Coronado, B.S. Tsukerblat, Mixed-
tion of the protonation of monomeric molybdic acid in sodium perchlorate valence polyoxometalate clusters. I. Delocalization of electronic pairs in
medium, J. Inorg. Nucl. Chem. 38 (1976) 2033e2036. dodecanuclear heteropoly blues with Keggin structure, Chem. Phys. 195
[18] J.J. Cruywagen, E.F.C.H. Rohwer, Thermodynamic constants for the first and (1995) 1e15.
second protonation of molybdate and the coordination of molybdenum(VI) [48] C. Sanchez, J. Livage, J.P. Launay, M. Fournier, Y. Jeannin, Electron delocal-
in the resulting monomeric species, J. S. Afr. Chem. I. 29 (1976) 30e39. ization in mixed-valence molybdenum polyanions, J. Am. Chem. Soc. 104
[19] J.J. Cruywagen, E.F.C.H. Rohwer, Coordination number of molybdenum(VI) in (1982) 3194e3202.
monomeric molybdic acid, Inorg. Chem. 14 (1975) 3136e3137. [49] S.D. Katewa, S.S. Katyare, A simplified method for inorganic phosphate
[20] M. Takahashi, Y. Abe, M. Tanaka, Elucidation of molybdosilicate complexes in determination and its application for phosphate analysis in enzyme assays,
the molybdate yellow method by ESI-MS, Talanta 131 (2015) 301e308. Anal. Biochem. 323 (2003) 180e187.
[21] N. Mahadevaiah, B. Venkataramani, B.S. Jai Prakash, Restrictive entry of [50] J.D. Burton, J.P. Riley, Determination of soluble phosphate, and total phos-
aqueous molybdate species into surfactant modified montmorillonite: a phorus in sea water and of total phosphorus in marine muds, Mikrochim.
breakthrough curve study, Chem. Mater. 19 (2007) 4606e4612. Acta (1956) 1350e1365.
[22] J.J. Cruywagen, Protonation, oligomerization, and condensation reactions of [51] T.G. Towns, Determination of aqueous phosphate by ascorbic acid reduction
vanadate(V), molybdate(vi), and tungstate(vi), in: A.G. Sykes (Ed.), Advances of phosphomolybdic acid, Anal. Chem. 58 (1986) 223e229.
in Inorganic Chemistry, Academic Press, 1999, pp. 127e182. [52] O. Broberg, K. Pettersson, Analytical determination of orthophosphate in
[23] C.C. Kircher, S.R. Crouch, Kinetics of the formation and decomposition of 12- water, Hydrobiologia 170 (1988) 45e59.
molybdophosphate, Anal. Chem. 55 (1983) 242e248. [53] S. Patachia, L. Isac, A comparative study of the spectrophotometric methods
80 E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82

used for phosphorus determination (II), Bull. Transilv. Univ. Brasov Ser. B 5 Naturwissenschaften 49 (1962) 513e514.
(1999) 53e58. [88] X. Lo pez, J.M. Maestre, C. Bo, J.-M. Poblet, Electronic properties of poly-
[54] H.-R. Sun, S.-Y. Zhang, J.-Q. Xu, G.-Y. Yang, T.-S. Shi, Electrochemical and in- oxometalates: a DFT study of a/b-[XM12O40]n- relative stability (M ¼ W, Mo
situ UV-visible-near-IR and FTIR spectroelectrochemical characterisation of and X a main group element), J. Am. Chem. Soc. 123 (2001) 9571e9576.
the mixed-valence heteropolyanion PMo12On 40 (n¼4, 5, 6, 7) in aprotic me- [89] T. Ueda, K. Isai, Effects of organic solvents and salts on the isomerization
dia, J. Electroanal. Chem. 455 (1998) 57e68. reaction (b->a) of Keggin-type 12-Molybdophosphate anion, Anal. Sci. 29
[55] J.M. Fruchart, G. Herve, J.P. Launay, R. Massart, Electronic spectra of mixed (2013) 447e453.
valence reduced heteropolyanions, J. Inorg. Nucl. Chem. 38 (1976) [90] V.W. Truesdale, P.J. Smith, C.J. Smith, Kinetics of alpha- and beta-
1627e1634. molybdosilicic acid formation, Analyst 104 (1979) 897e918.
[56] I. Koshiishi, T. Imanari, Study of the coloured substances in molybdenum [91] V.W. Truesdale, C.J. Smith, The automatic determination of silicate dissolved
blue using high-performance liquid chromatography, J. Chromatogr. A 358 in natural fresh water by means of procedures involving the use of either a-
(1986) 195e200. or b-molybdosilicic acid, Analyst 101 (1976) 19e31.
[57] E. Papaconstantinou, Photochemistry of polyoxometallates of molybdenum [92] V.W. Truesdale, C.J. Smith, P.J. Smith, Transformation and decomposition of
and tungsten and/or vanadium, Chem. Soc. Rev. 18 (1989) 1e31. b-molybdosilicic acid, Analyst 102 (1977) 73e85.
[58] R. Massart, First stages in the reduction of B-molybdosilicic acid, Ann. Chim. [93] R.A. Chalmers, A.G. Sinclair, Analytical applications of b-heteropoly acids:
Paris 4 (1969) 365e370. part I. Determination of arsenic, germanium and silicon, Anal. Chim. Acta 33
[59] S.J. Eisenreich, J.E. Going, Extraction of reduced molybdophosphoric and (1965) 384e390.
molybdoantimonylphosphoric acids with oxygenated solvents, Anal. Chim. [94] J.P. Launay, R. Massart, P. Souchay, Gradual reduction of molybdosilicates and
Acta 71 (1974) 393e403. related compounds, J. Less Common Met. 36 (1974) 139e150.
[60] F. Huey, L.G. Hargis, Spectrophotometric determination of cesium using 12- [95] J. Steigman, W.C. Eckelman, The Chemistry of Technetium in Medicine, Na-
molybdophosphoric acid, Anal. Chem. 39 (1967) 125e127. tional Academy Press, Washington, D. C., 1992.
[61] R. Contant, 12-Arseno-b-molybdic acid, C. R. Acad. Sci. 267 (1968) 1479. [96] R. Massart, M. Fournier, P. Souchay, Effect of stannous chloride on molyb-
[62] T.D. Fontaine, Spectrophotometric determination of phosphorus, Ind. Eng. dosilicic acid: detection of reduced molybdostannosilicates, C. R. Acad. Sci.
Chem. Anal. Ed. 14 (1942) 77e78. 267 (1968) 1805e1808.
[63] R.P.A. Sims, Formation of heteropoly blue by some reduction procedures [97] E.A. Nagul, I.D. McKelvie, S.D. Kolev, The nature of the salt error in the Sn(II)-
used in the micro-determination of phosphorus, Analyst 86 (1961) 584e590. reduced molybdenum blue reaction for determination of dissolved reactive
[64] S.R. Dickman, R.H. Bray, Colorimetric determination of phosphate, Ind. Eng. phosphorus in saline waters, Anal. Chim. Acta (2015) (under revision).
Chem. Anal. Ed. 12 (1940) 665e668. [98] APHA, Standard Methods for the Examination of Water & Wastewater,
[65] A.-A. El Sayed, Y. Hussein, M. Mohammed, Simultaneous determination of AWWA, 1999, pp. 249e254.
phosphate and silicate by first-derivative spectrophotometry, Anal. Sci. 17 [99] E.A. Nagul, C. Fonta s, I.D. McKelvie, R.W. Cattrall, S.D. Kolev, The use of a
(2001) 1461e1464. polymer inclusion membrane for on-line separation and preconcentration of
[66] E.E. Kriss, V.K. Rudenko, K.B. Yatsimirskii, Reduction of molybdate to mo- dissolved reactive phosphorus, Anal. Chim. Acta 803 (2013) 82e90.
lybdenum blue by various reducing agents, Zh. Neorg. Chim. 16 (1971) [100] H. Levine, J.J. Rowe, F.S. Grimaldi, The molybdenum blue reaction and the
2147e2153. determination of phosphorus in waters containing arsenic, silicon and
[67] H. Levine, J.J. Rowe, F.S. Grimaldi, Molybdenum blue reaction and determi- germanium, Science 119 (1954) 327e328.
nation of phosphorus in waters containing arsenic, silicon, and germanium, [101] A. Sjo€ sten, S. Blomqvist, Influence of phosphate concentration and reaction
Anal. Chem. 27 (1955) 258e262. temperature when using the molybdenum blue method for determination of
[68] G. Hesse, K. Geller, Phosphorus (nucleotide phosphorus) determination with phosphate in water, Water Res. 31 (1997) 1818e1823.
a stannous chloride hydrazine sulfate reagent by the Hurst method, Mikro- [102] H. Ishii, Determination of and coloring method and agent for phosphate ion
chim. Acta (1968) 526e533. in sample water by formation of molybdenum blue, in: J.P. Office (Ed.), Jpn.
[69] T.J. Han, An improved phosphorus assay for oils without carcinogenic hy- Kokai Tokkyo Koho, Miura Co., Japan, 2012, p. 11.
drazine sulfate, J. Am. Oil Chem. Soc. 72 (1995) 881e885. [103] R. Ammon, K. Hinsberg, Colorimetric determination of phosphoric and
[70] P.S. Chen, T.Y. Toribara, H. Warner, Microdetermination of phosphorus, Anal. arsenic acids with ascorbic acid, Z. Physiol. Chem. 239 (1936) 207e216.
Chem. 28 (1956) 1756e1758. [104] H. Kobayashi, E. Nakamura, Spectrophotometric determination of phos-
[71] O.H. Lowry, J.A. Lopez, The determination of inorganic phosphate in the phorus based on the formation of phosphomolybdenum blue by reduction
presence of labile phosphate esters, J. Biol. Chem. 162 (1946) 421e428. with bismuth (III)-ascorbic acid, Bunseki Kagaku 56 (2007) 561e566.
[72] S. Ganesh, F. Khan, M.K. Ahmed, P. Velavendan, N.K. Pandey, U.K. Mudali, [105] H. Kobayashi, E. Nakamura, Mechanism of salt error on the color develop-
Spectrophotometric determination of trace amounts of phosphate in water ment of phosphomolybdenum blue, Bunseki Kagaku 53 (2004) 119e122.
and soil, Water Sci. Technol. 66 (2012) 2653e2658. [106] M. Takahashi, M. Tanaka, Analysis of complex-formation reaction in mo-
[73] D.F. Boltz, M.G. Mellon, Determination of phosphorus, germanium, silicon lybdenum blue method by ESI-MS, Bunseki Kagaku 61 (2012) 1049e1054.
and arsenic by the heteropoly blue method, Anal. Chem. 19 (1947) 873e877. [107] F. Cavani, R. Mezzogori, A. Pigamo, F. Trifiro  , Improved catalytic performance
[74] W.D. Harris, P. Harris, Popat, Determination of the phosphorus content of of Keggin-type polyoxometalates in the oxidation of isobutane to meth-
lipids, J. Am. Oil Chem. Soc. 31 (1954) 124e127. acrylic acid under hydrocarbon-lean conditions using antimony-doped cat-
[75] B.L. Griswold, F.L. Humoller, A.R. McIntyre, Inorganic phosphates and phos- alysts, Chem. Eng. J. 82 (2001) 33e42.
phate esters in tissue extracts, Anal. Chem. 23 (1951) 192e194. [108] F. Cavani, A. Tanguy, F. Trifiro  , M. Koutrev, Effect of antimony on the
[76] H. Weil-Malherbe, The catalytic effect of molybdate on the hydrolysis of chemicalephysical features and reactivity in isobutyric acid oxidehydroge-
organic phosphates, Biochem. J. 55 (1953) 741e745. nation of Keggin-type heteropolycompounds, J. Catal. 174 (1998) 231e241.
[77] C.H. Fiske, Y. Subbarow, The colorimetric determination of phosphorus, [109] J.Z. Zhang, C.J. Fischer, P.B. Ortner, Optimization of performance and mini-
J. Biol. Chem. 66 (1925) 375e400. mization of silicate interference in continuous flow phosphate analysis,
[78] J. Huo, Q. Li, Determination of thiamazole in pharmaceutical samples by Talanta 49 (1999) 293e304.
phosphorus molybdenum blue spectrophotometry, Spectrochim. Acta, Part A [110] L. Dermeche, R. Thouvenot, S. Hocine, C. Rabia, Preparation and character-
87 (2012) 293e297. ization of mixed ammonium salts of Keggin phosphomolybdate, Inorg. Chim.
[79] F.B. Salem, Spectrophotometric determination of phosphate in waters of Acta 362 (2009) 3896e3900.
Egypt, Water Air Soil Pollut. 60 (1991) 27e33. [111] X.-L. Huang, J.-Z. Zhang, Kinetic spectrophotometric determination of sub-
[80] P.K. Gupta, R. Ramachandran, Spectrophotometric determination of phos- micromolar orthophosphate by molybdate reduction, Microchem. J. 89
phorus in steel using phosphoantimonyl molybdate complex, Microchem. J. (2008) 58e71.
26 (1981) 32e39. [112] S.-C. Pai, T.-Y. Wang, T.-H. Fang, K.-T. Jiann, Effect of heating on the color
[81] J.E. Going, S.J. Eisenreich, Spectrophotometric studies of reduced molyb- formation reaction in the Murphy and Riley method for the determination of
doantimonylphosphoric acid, Anal. Chim. Acta 70 (1974) 95e106. phosphate in natural waters, J. Environ. Anal. Chem. 2 (2015) 139.
[82] S.-C. Pai, C.-C. Yang, J.P. Riley, Effects of acidity and molybdate concentration [113] E.A.G. Zagatto, C.C. Oliveira, A. Townshend, P.J. Worsfold, Flow Analysis with
on the kinetics of the formation of the phosphoantimonylmolybdenum blue Spectrophotometric and Luminometric Detection, Elsevier, Amsterdam,
complex, Anal. Chim. Acta 229 (1990) 115e120. 2012.
[83] J.E. Harwood, R.A. van Steenderen, A.L. Kühn, A rapid method for ortho- [114] V.W. Truesdale, C.J. Smith, Formation of molybdosilicic acids from mixed
phosphate analysis at high concentrations in water, Water Res. 3 (1969) solutions of molybdate and silicate, Analyst 100 (1975) 203e212.
417e423. [115] J.D.H. Strickland, The preparation and properties of silicomolybdic acid. III.
[84] G.P. Edwards, A.H. Molof, R.W. Schneeman, Determination of orthophos- The combination of silicate and molybdate, J. Am. Chem. Soc. 74 (1952)
phate in fresh and saline waters, J. Am. Water Works Assoc. 57 (1965) 872e876.
917e925. [116] CRC Handbook of Chemistry & Physics, 95 ed., CRC Press, 2014e2015.
[85] J. Murphy, J.P. Riley, A modified single solution method for the determination [117] H.W. Harvey, The estimation of phosphate and of total phosphorus in sea
of phosphate in natural waters, Anal. Chim. Acta 27 (1962) 31e36. waters, J. Mar. Biol. Assoc. U. K. 27 (1948) 337e359.
[86] J.N. Barrows, G.B. Jameson, M.T. Pope, Structure of a heteropoly blue. The [118] B. Botar, A. Ellern, P. Ko €gerler, Mapping the formation areas of giant mo-
four-electron reduced.beta.-12-molybdophosphate anion, J. Am. Chem. Soc. lybdenum blue clusters: a spectroscopic study, J. Chem. Soc. Dalton Trans. 41
107 (1985) 1771e1773. (2012) 8951.
[87] H. Hahn, W. Becker, Polarographic behavior of molybdenum heteropoly [119] T. Akutagawa, R. Jin, R. Tunashima, S. Noro, L. Cronin, T. Nakamura, Nano-
acids. VIII. Two reduced forms of molybdophosphoric acid and their salts, scale assemblies of gigantic molecular {Mo154}-rings:
E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82 81

(dimethyldioctadecylammonium)20[Mo154O462H8(H2O)70], Langmuir 24 [149] I.D. McKelvie, B.T. Hart, T.J. Cardwell, R.W. Cattrall, Spectrophotometric
(2008) 231e238. determination of dissolved organic phosphorus in natural waters using
[120] A. Müller, J. Meyer, E. Krickemeyer, E. Diemann, Molybdenum blue: a 200 in-line photooxidation and flow injection, Analyst 114 (1989)
year old mystery unveiled, Angew. Chem. Int. Ed. Engl. 35 (1996) 1459e1463.
1206e1208. [150] D.M.W. Peat, I.D. McKelvie, G.P. Matthews, P.M. Haygarth, P.J. Worsfold,
[121] A. Komura, Y. Ikeda, H. Imanaga, Reactions of the Di-m-oxo-bis[aqua- Rapid determination of dissolved organic phosphorus in soil leachates and
oxalatooxomolybdate(V)] ion, Bull. Chem. Soc. Jpn. 49 (1976) 131e137. runoff waters by flow injection analysis with on-line photo-oxidation,
[122] M. Ardon, A. Pernick, Molybdenum(V) in aqueous solution, Inorg. Chem. 12 Talanta 45 (1997) 47e55.
(1973) 2484e2485. [151] E.S. Baginski, P.P. Foa, B. Zak, Determination of phosphate: study of labile
[123] L.H.N. Cooper, Salt error in determinations of phosphate in sea water, J. Mar. organic phosphate interference, Clin. Chim. Acta 15 (1967) 155e158.
Biol. Assoc. U. K. 23 (1938) 171e178. [152] D.S. Baldwin, Reactive “organic” phosphorus revisited, Water Res. 32 (1998)
[124] J.T. Woods, M.G. Mellon, The molybdenum blue reaction e A spectropho- 2265e2270.
tometric study, Ind. Eng. Chem. 13 (1941) 760. [153] X.-L. Huang, J.-Z. Zhang, Neutral persulfate digestion at sub-boiling tem-
[125] E.A. Nagul, I.D. McKelvie, S.D. Kolev, The use of on-line UV photoreduction in perature in an oven for total dissolved phosphorus determination in natural
the flow analysis determination of dissolved reactive phosphate in natural waters, Talanta 78 (2009) 1129e1135.
waters, Talanta 133 (2015) 155e161. [154] F. Lipmann, L.C. Tuttle, Acetyl phosphate: chemistry, determination, and
[126] T. Rupasinghe, T.J. Cardwell, R.W. Cattrall, M.D.L. de Castro, S.D. Kolev, Per- synthesis, J. Biol. Chem. 153 (1944) 571e582.
vaporation-flow injection determination of arsenic based on hydride gen- [155] C.H. Fiske, Y. Subbarow, Phosphocreatine, J. Biol. Chem. 81 (1929) 629e679.
eration and the molybdenum blue reaction, Anal. Chim. Acta 445 (2001) [156] S. Tarapchak, Soluble reactive phosphorus measurements in lake water:
229e238. evidence for molybdate-enhanced hydrolysis, J. Environ. Qual. 12 (1983)
[127] C. Janardanan, S.M. Nair, Studies on inorganic ion exchangers. Part 5. Prep- 105.
aration, properties and application of antimony(III) arsenate and anti- [157] A.V.S.V. Cavaleiro, V.S. Gil, J. Pedrosa de Jesus, R. Gillard, P. Williams, N.m.r.
mony(III) molybdate, Analyst 115 (1990) 85e87. studies of complexes of molybdenum(VI) with tartaric acid in aqueous so-
[128] J.-Z. Zhang, L. Guo, C. Fischer, Abundance and chemical speciation of phos- lution, Transit. Met. Chem. 9 (1984) 62e67.
phorus in sediments of the Mackenzie River Delta, the Chukchi Sea and the [158] J.J. Cruywagen, J.B.B. Heyns, E.A. Rohwer, Molybdenum(VI) complex forma-
Bering Sea: importance of detrital apatite, Aquat. Geochem. 16 (2010) 353e371. tion. Part 4. Equilibria and thermodynamic quantities for the reactions with
[129] J.M. Coddington, M.J. Taylor, Molybdenum-95 nuclear magnetic resonance tartrate in 1.0 mol dm3 sodium chloride, J. Chem. Soc. Dalton Trans. (1990)
and vibrational spectroscopic studies of molybdenum(VI) species in aqueous 1951e1956.
solutions and solvent extracts from hydrochloric and hydrobromic acid: [159] M. Cindri c, N. Strukan, V. Vrdoljak, M. Dev ci
c, Z. Veksli, B. Kamenar, Syn-
evidence for the complexes [Mo2O5(H2O)6]2þ, [MoO2X2(H2O)2] (X ¼ Cl or Br), thesis, structure and properties of molybdenum(VI) oxalate complexes of the
and [MoO2Cl4]2, J. Chem. Soc. Dalton Trans. (1990) 41e47. types M2[Mo2O5(C2O4)2(H2O)2] and M2[MoO3(C2O4)] (M¼Na, K, Rb, Cs),
[130] G.C. Dehne, M.G. Mellon, Spectrophotometric estimation of zirconium as Inorg. Chim. Acta 304 (2000) 260e267.
reduced molybdosulfatozirconic acid, Anal. Chem. 35 (1963) 1382e1386. [160] M. Cindri c, N. Strukan, V. Vrdoljak, T. Fuss, G. Giester, B. Kamenar, Synthesis
[131] T. Osakai, S. Himeno, A. Saito, T. Hori, Voltammetric determination of sul- and structures of ammonium and tetraphenylphosphonium salts of e -oxo-
phate ion through heteropoly blue formation, J. Electroanal. Chem. 278 diaquadioxalatotetraoxodimolybdenum(VI). An interesting example of
(1990) 217e225. intramolecular hydrogen bonds within the dimeric anion, Inorg. Chim. Acta
[132] K.G. Falk, K. Sugiura, Precipitation of phosphorus as ammonium phospho- 309 (2000) 77e81.
molybdate in the presence of sulfuric acid, J. Am. Chem. Soc. 37 (1915) [161] J.J. Cruywagen, J.B.B. Heyns, R.F. van de Water, A potentiometric, spectro-
1507e1515. photometric, and calorimetric investigation of molybdenum(VI)-oxalate
[133] S. Himeno, Y. Ueda, M. Hasegawa, Spectrophotometric investigation on the complex formation, J. Chem. Soc. Dalton Trans. (1986) 1857e1862.
equilibration of monomeric forms of Mo(VI) in aqueous sulfuric acid, Inorg. [162] J.J. Cruywagen, E.A. Rohwer, R.F. van de Water, Molybdenum(VI) complex
Chim. Acta 70 (1983) 53e57. formation. Equilibria and thermodynamic quantities for the reactions with
[134] P. Linares, L. De Castro, M. Valcarcel, Flow injection analysis of binary and malate, Polyhedron 16 (1997) 243e251.
ternary mixtures of arsenite, arsenate, and phosphate, Anal. Chem. 58 (1986) [163] Z.-H. Zhou, Y.-F. Deng, Z.-X. Cao, R.-H. Zhang, Y.L. Chow, Dimeric dioxomo-
120e124. lybdenum(VI) and oxomolybdenum(V) complexes with citrate at very low
[135] M. Lopez Carreto, D. Sicilia, S. Rubio, D. Perez-Bendito, Simultaneous deter- pH and neutral conditions, Inorg. Chem. 44 (2005) 6912e6914.
mination of arsenate and phosphate by use of the kinetic wavelength-pair [164] R.-H. Zhang, X.-W. Zhou, Y.-C. Guo, M.-L. Chen, Z.-X. Cao, Y.L. Chow, Z.-
method, Anal. Chim. Acta 283 (1993) 481e488. H. Zhou, Crystalline and solution chemistry of tetrameric and dimeric
[136] S. Hu, J. Lu, C. Jing, A novel colorimetric method for field arsenic speciation molybdenum(VI) citrato complexes, Inorg. Chim. Acta 406 (2013) 27e36.
analysis, J. Env. Sci. 24 (2012) 1341e1346. [165] L.L. Wei, C.R. Chen, Z.H. Xu, The effect of low-molecular-weight organic acids
[137] D.L. Johnson, M.E.Q. Pilson, Spectrophotometric determination of arsenite, and inorganic phosphorus concentration on the determination of soil
arsenate, and phosphate in natural waters, Anal. Chim. Acta 58 (1972) phosphorus by the molybdenum blue reaction, Biol. Fert. Soils 45 (2009)
289e299. 775e779.
[138] S. Tsang, F. Phu, M.M. Baum, G.A. Poskrebyshev, Determination of phosphate/ [166] B. Raben-Lange, A.B. Bendtsen, S.S. Jørgensen, Spectrophotometric determi-
arsenate by a modified molybdenum blue method and reduction of arsenate nation of silicon in soil solutions by flow injection analysis: reduction of
by S2O24 , Talanta 71 (2007) 1560e1568. phosphate interference, Commun. Soil Sci. Plant Anal. 25 (1994) 3241e3256.
[139] R.E. Stauffer, Determination of arsenic and phosphorus compounds in [167] F.P. Sudakov, L.V. Butorova, Effect of complex-forming substances on the
groundwater with reduced molybdenum blue, Anal. Chem. 55 (1983) reduction of molybdophosphoric acid, Zh. Anal. Khim. 23 (1968) 721e726.
1205e1210. [168] Q. Zini, P.L. Buldini, L. Morettini, Rapid determination of dissolved silica in
[140] R.K. Dhar, Y. Zheng, J. Rubenstone, A. van Geen, A rapid colorimetric method natural waters, Microchem. J. 32 (1985) 148e152.
for measuring arsenic concentrations in groundwater, Anal. Chim. Acta 526 [169] C. Bergamini, Masking of molybdenum complexes by fluoride ions. Spec-
(2004) 203e209. trophotometric study, Anal. Chim. Acta 4 (1950) 153e158.
[141] J.B. Mullin, J.P. Riley, The colorimetric determination of silicate with special [170] R.H. Campbell, M.G. Mellon, An indirect absorptiometric method for the
reference to sea and natural waters, Anal. Chim. Acta 12 (1955) 162e176. determination of boron, Anal. Chem. 32 (1960) 50e54.
[142] E. Weitz, Silicic acid and silicates, Chem. Ztg. 74 (1950) 256e257. [171] E.N. Kryachko, Y.V. Karyakin, Effect of fluoride ion on molybdic acid poly-
[143] F.R. Campbell, R.L. Thomas, Automated method for determining and merization, Zh. Neorg. Chim. 15 (1970) 26.
removing silica interference in determination of soluble phosphorus in lake [172] S. Blomqvist, K. Hjellstro € m, A. Sjo€sten, Interference from arsenate, fluoride
and stream waters, Environ. Sci. Technol. 4 (1970) 602e604. and silicate when determining phosphate in water by the phosphoantimo-
[144] R.A. Chalmers, A.G. Sinclair, Analytical applications of b-heteropoly acids: nylmolybdenum blue method, Intern. J. Environ. Anal. Chem. 54 (1993)
part II. The influence of complexing agents on selective formation, Anal. 31e43.
Chim. Acta 34 (1966) 412e418. [173] L.T. Kurtz, Elimination of fluoride interference in the molybdenum blue re-
[145] C.X. Galhardo, J.C. Masini, Spectrophotometric determination of phosphate action, Ind. Eng. Chem. Anal. Ed. 14 (1942) 855.
and silicate by sequential injection using molybdenum blue chemistry, Anal. [174] O.P. Bhargava, G.F. Pitt, W.G. Hines, Unified automated determination of
Chim. Acta 417 (2000) 191e200. silicon in iron ores, sinters, slags, iron and steel, Talanta 18 (1971) 793e798.
[146] L.J. Gimbert, P.M. Haygarth, P.J. Worsfold, Determination of nanomolar [175] N. Zhou, Mechanism of sodium fluoride-tin(II) chloride solution as reducing
concentrations of phosphate in natural waters using flow injection with a agent, Lihua Jianyan, Huaxue Fence 28 (1992) 239.
long path length liquid waveguide capillary cell and solid-state spectro- [176] J. Murphy, J.P. Riley, A single-solution method for the determination of sol-
photometric detection, Talanta 71 (2007) 1624e1628. uble phosphate in sea water, J. Mar. Biol. Assoc. U. K. 37 (1958) 9e14.
[147] Z.L. He, V.C. Baligar, K.D. Ritchey, D.C. Martens, Determination of soluble [177] B. Müller, T.M. Seward, Spectrophotometric determination of the stability of
phosphorus in the presence of organic ligands or fluoride, Soil Sci. Soc. Am. J. tin(II) chloride complexes in aqueous solution up to 300 C, Geochim. Cos-
62 (1998) 1538e1541. mochim. Acta 65 (2001) 4187e4199.
[148] R.L. Benson, I.D. McKelvie, B.T. Hart, Y.B. Truong, I.C. Hamilton, Determina- [178] T. Gajda, P. Sipos, H. Gamsja €ger, The standard electrode potential of
tion of total phosphorus in waters and wastewaters by on-line UV/thermal the Sn4þ/Sn2þ couple revisited, Monatsh. Chem. 140 (2009)
induced digestion and flow injection analysis, Anal. Chim. Acta 326 (1996) 1293e1303.
29e39. [179] S.W. Rabideau, R.H. Moore, The application of high-speed computers to the
82 E.A. Nagul et al. / Analytica Chimica Acta 890 (2015) 60e82

least squares determination of the formation constants of the chloro- dissolved reactive phosphorus in estuarine waters using a reversed flow
complexes of tin(II), J. Phys. Chem. 65 (1961) 371e373. injection manifold, Analyst 122 (1997) 1477e1480.
[180] I.D. McKelvie, D.M.W. Peat, G.P. Matthews, P.J. Worsfold, Elimination of [204] J.-Z. Zhang, Distinction and quantification of carry-over and sample inter-
the Schlieren effect in the determination of reactive phosphorus in action in gas segmented continuous flow analysis, J. Autom. Chem. 19
estuarine waters by flow-injection analysis, Anal. Chim. Acta 351 (1997) (1997).
265e271. [205] A. Cifuentes, J.L. Bernal, J.C. Diez-Masa, Determination of critical micelle
[181] G. Nürnberg, Iron and hydrogen sulfide interference in the analysis of soluble concentration values using capillary electrophoresis instrumentation, Anal.
reactive phosphorus in anoxic waters, Water Res. 18 (1984) 369e377. Chem. 69 (1997) 4271e4274.
[182] N.A. Zatar, M.A. Abu-Eid, A.F. Eid, Spectrophotometric determination of ni-
trite and nitrate using phosphomolybdenum blue complex, Talanta 50
(1999) 819e826. Edward A. Nagul is a PhD student in the School of
[183] M. Grace, Y. Udnan, I. McKelvie, J. Jakmunee, K. Grudpan, On-line removal of Chemistry at the University of Melbourne. His research
sulfide interference in phosphate determination by flow injection analysis, focuses on developing flow injection methodology for
Environ. Chem. 3 (2006) 19e25. ultra-trace nutrient analysis in terrestrial and marine wa-
[184] V.N. De Jonge, L.A. Villerius, Interference of sulphide in inorganic phosphate ters, and on the development of polymer inclusion mem-
determination in natural waters, Mar. Chem. 9 (1980) 191e197. branes for the efficient separation and preconcentration
[185] P.L. Buldini, P. Saxena, A. Toponi, Q. Zini, Determination of phosphate and of these analytes. Of particular interest has been the use
silicate in lead-acid electrolyte using molybdenum blue, Microchem. J. 38 of on-line photochemical reduction in orthophosphate
(1988) 399e402. determination, and the use of fluorescence microscopy
[186] M.Z. Barakat, S.K. Shehab, The microdetermination of ferrous iron, Micro- and small-angle X-ray scattering in elucidating the inter-
chem. J. 8 (1964) 6e11. nal structure of polymer inclusion membranes.
[187] C. Wang, T. Hu, H. Ho, Two types of phosphorus molybdenum blue in
iron(III)-tin(II)-fluoride reducing systems, Chin. J. Anal. Chem. 6 (1978)
83.
[188] L. Duval, Influence of silicon, germanium, and iron(III) on the molyb-
denum blue determination of phosphoric acid, Chim. Anal. Paris 48
Ian D. McKelvie is Principal Fellow in the School of
(1966) 290.
Chemistry at the University of Melbourne and Visiting
[189] S. Fan, Z. Fang, Compensation of calibration graph curvature and interference
Professor, School of Geography, Earth, and Environmental
in flow-injection spectrophotometry using gradient ratio calibration, Anal.
Sciences, at the Plymouth University, UK. Prior to that he
Chim. Acta 241 (1990) 15e22.
was Associate Professor at Monash University, where his
[190] J. Malý, Phosphorus determination in sewage sludge, Acta Hydrochim.
research focused on the development of flow-based
Hydrobiol. 13 (1985) 137e147.
analytical systems for water quality assessment of macro-
[191] N. Ichinose, H. Kanai, K. Nakamura, C. Shimizu, H. Kurokura, K. Okamoto,
nutrients in marine, estuarine and freshwater systems, and
T. Inui, A problem in the spectrophotometric determination of dissolved
on the biogeochemistry of organic phosphates. He has co-
total phosphorus in brackish anoxic waters, Anal. Chim. Acta 156 (1984)
edited two monographs: Environmental Monitoring Hand-
345e349.
book (2002) and Advances in flow injection analysis and
[192] J. Liu, Problem in determination of silicon dioxide content in iron concen-
related techniques (2008), and joined the journal Talanta
trate powder by molybdenum blue spectrophotometry and method
as an associate editor in 2013.
improvement thereof, Dangdai Huagong 42 (2013) 709e710.
[193] P. Pakalns, H. Pakalns, Steman, Effect of surfactants on the spectrophoto-
metric determination of phosphate by direct and extraction procedures,
Water Res. 10 (1976) 437e441. Paul J. Worsfold is Professor of Analytical Chemistry at
[194] J.M. Estela, V. Cerd a, Flow analysis techniques for phosphorus: an overview, Plymouth University (since 1990) and Co-Director of the
Talanta 66 (2005) 307e331. Biogeochemistry Research Centre. He obtained his BSc at
[195] J.-Z. Zhang, C.J. Fischer, P.B. Ortner, Continuous flow analysis of phosphate in Loughborough University (1976) and his PhD at the Uni-
natural waters using hydrazine as a reductant, Intern. J. Environ. Anal. Chem. versity of Toronto (1980). He is Chair of the Division of
80 (2001) 61e73. Analytical Chemistry of the European Association for
[196] Y. Hirai, N. Yoza, S. Ohashi, Flow injection analysis of inorganic ortho- and Chemical and Molecular Sciences and a Past-President of
polyphosphates using ascorbic acid as a reductant of molybdophosphate, the Analytical Division of the Royal Society of Chemistry.
Chem. Lett. (1980) 499e502. One research interest is techniques for the determination
[197] S. Somnam, K. Grudpan, J. Jakmunee, Stopped-flow injection method for of macronutrients and micronutrients and understanding
determination of phosphate in soils and fertilisers, Maejo Int. J. Sci. Tech. 2 their role in terrestrial and aquatic biogeochemical pro-
(2008) 172e181. cesses. Flow injection provides an integrating theme for
[198] Y. Narusawa, Separation and simultaneous determination of phosphate, this research and he has recently co-authored a textbook
arsenate, and silicate with on-line column flow injection analysis, J. Flow. Inj. on the subject.
Anal. 4 (1987) 20.
[199] L.J. Gimbert, P.J. Worsfold, Environmental applications of liquid-waveguide-
capillary cells coupled with spectroscopic detection, Trends Anal. Chem. 26 Spas D. Kolev completed his undergraduate degree in
(2007) 914e930. chemistry at the University of Sofia (Bulgaria) in 1982. In
[200] J.-Z. Zhang, Enhanced sensitivity in flow injection analysis using a long 1988 he obtained his PhD in the area of analytical chem-
pathlength liquid waveguide capillary flow cell for spectrophotometric istry from the Budapest University of Technology and
detection, Anal. Sci. 22 (2006) 57e60. Economics (Hungary). After working as an academic at the
[201] Q.P. Li, D.A. Hansell, Intercomparison and coupling of magnesium-induced University of Sofia, University of Twente (The Netherlands)
co-precipitation and long-path liquid-waveguide capillary cell techniques and La Trobe University (Australia), Spas joined the School
for trace analysis of phosphate in seawater, Anal. Chim. Acta 611 (2008) of Chemistry of The University of Melbourne in 2001
68e72. where he is now Professor of Chemistry. One of his main
[202] J.-Z. Zhang, J. Chi, Automated analysis of nanomolar concentrations of research areas is chemical analysis where he has devel-
phosphate in natural waters with liquid waveguide, Environ. Sci. Technol. 36 oped numerous flow analysis methods and associated
(2002) 1048e1053. equipment, microfluidic paper-based analytical devices,
[203] S. Auflitsch, D.M.W. Peat, P.J. Worsfold, I.D. McKelvie, Determination of and chemical sensors.

Vous aimerez peut-être aussi